HARVARD UNIVERSITY Library of the Museum of Comparative Zoology SREAT BASIN NATURALIST MEMOIR: nber 2 Brigham Young University JUN 16 1978 d*\ HARVARD INTERMOUNTAIN w££m BIOGEOGRAPHY: ^v W A SYMPOSIUM 197 GREAT BASIN NATURALIST MEMOIRS Editor. Stephen L. Wood, Department of Zoology, Brigham Young University, Provo, Utah 84602. Editorial Board. Kimball T. Harper, Botany; Wilmer W. Tanner, Life Science Museum; Stanley L. Welsh, Botany; Clayton M. White, Zoology. Ex Officio Editorial Board Members. A. Lester Allen, dean, College of Biological and Agri- cultural Sciences; Ernest L. Olson, director, Brigham Young University Press, Univer- sity Editor. The Great Basin Naturalist was founded in 1939 by Vasco M. Tanner. It has been published from one to four times a year since then by Brigham Young University, Provo, Utah. In general, only previously unpublished manuscripts of less than 100 printed pages in length and pertaining to the biological natural history of western North America are ac- cepted. The Great Basin Naturalist Memoirs was established in 1976 for scholarly works in biological natural history longer than can be accommodated in the parent publication. The Memoirs appears irregularly and bears no geographical restriction in subject matter. Manu- scripts are subject to the approval of the editor. Subscriptions. The annual subscription to the Great Basin Naturalist is $12 (outside the United States $13). The price for single numbers is $4 each. All back numbers are in print and are available for sale. All matters pertaining to the purchase of subscriptions and back numbers should be directed to Brigham Young University, Life Science Museum, Pro- vo, Utah 84602. The Great Basin Naturalist Memoirs may be purchased from the same of- fice at the rate indicated on the inside of the back cover of either journal. Scholarly Exchanges. Libraries or other organizations interested in obtaining either journal through a continuing exchange of scholarly publications should contact the Brigham Young University Exchange Librarian, Harold B. Lee Library, Provo, Utah 84602. Manuscripts. All manuscripts and other copy for either the Great Basin Naturalist or the Great Basin Naturalist Memoirs should be addressed to the editor as instructed on the back cover. RE AT BASIN NATURALIST MEMOIR! INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM Printed in the United States of America by Brigham Young University Printing Service 3-78 1.5M 29246 REAT BASIN NATURALIST MEMOIRS Number 2 Brigham Young University 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM K. T. Harper and James L. Reveal Symposium Organizers CONTENTS Page Preface 1 The biota of the Intermountain Region in geohistorical context. Arthur Cronquist .... 3 Biogeography of intermountain fishes. Gerald R. Smith 17 Zoogeography of reptiles and amphibians in the Intermountain Region. Wilmer W. Tanner 43 Avian biogeography of the Great Basin and Intermountain Region. William H. Behle 55 The flora of Great Basin mountain ranges: Diversity, sources, and dispersal ecology. K. T. Harper, D. C. Freeman, W. K. Ostler and L. G. Klikoff 81 Alpine phytogeography across the Great Basin. W. D. Billings 105 Phytogeographical variation within juniper-pinyon woodlands of the Great Basin. Neil E. West, Robin J. Tausch, Kenneth H. Rea, and Paul T. Tueller 119 Patterns of avian geography and speciation in the Intermountain Region. Ned K. Johnson 137 Explosive evolution of perennial Atriplex in western America. Howard C. Stutz 161 Distribution and phylogeny of Eriogonoideae (Polygonaceae). James L. Reveal 169 Problems in plant endemism on the Colorado Plateau. Stanley L. Welsh 191 Some factors governing plant distributions in the Mojave-Intermountain Transition Zone. Susan E. Meyer 197 The theory of insular biogeography and the distribution of boreal birds and mam- mals. James H. Brown 209 Biogeography and management of native western shrubs: A case study, Section Tri- dentatae of Artemisia. E. Durant Mc Arthur and A. Perry Plummer 229 Applying biogeographic principles to resource management: A case study evaluating Holdridge's Life Zone model. James A. MacMahon and Thomas F. Wieboldt ... 245 Index 259 No. 2 Great Basin Naturalist Memoirs Intermountain Biogeography: A Symposium Brigham Young University, Provo, Utah 1978 K. T. Harper' and James L. Reveal2 PREFACE Most of the Intermountain Region is re- mote from the nation's major transportation arterials. As a consequence, the region's beauty and its biological resources are largely . unappreciated. The area is often considered to be a wasteland, a desert with little or no life, and a land of minimal val- ue. Because few initially understood or ap- preciated the region, its resources have of- ten been abused by a variety of human activities ranging from military weapon testing to off-road vehicle travel. Addition- ally, ranchers and governmental land man- agers have had no precedents to guide their activities in the unique and often fragile ecosystems of the region. As a consequence, the biological landscape has been markedly altered by grazing, control of wildfires, and agronomic practices. With time and knowl- edge, concern for the natural landscape of the West has grown. National and state leg- islation now imposes strict guidelines on most new development activities and re- quires a balanced analysis of the full impact of major activities on the land. Even rare native species have advocates in high places and now enjoy some legal protection. Although knowledge has accumulated and management of the biological resources of the region has steadily improved, there is still much to do. New facts must be ac- cumulated and those now known must be digested, disseminated, and put to work in management by individuals and govern- ments. This volume brings together current information on a broad spectrum of the na- tive biota of the region. Most of the authors Consider the management implications of their findings. We hope the volume will in- form and assist resource managers charged with preserving the ecological health of the Intermountain West. For the biologist, this book presents the first major overview of current biogeogra- phical research being conducted by a num- ber of individuals and institutions in the in- termountain region. Of more importance, however, this symposium represents the first attempt to bring both plant and animal sci- entists and management-oriented researchers together to discuss and evaluate the bio- geographic principles at work in the area. Topics discussed include distribution pat- terns for fishes, reptiles, amphibians, birds, 'Department of Botany and Range Science, Brigham Young University, Provo, Utah 84602. 'Department of Botany, University of Maryland, College Park, Maryland 20742, and National Museum of Natural Hi! Washington, DC. 20560. Smithsonian Institution, GREAT BASIN NATURALIST MEMOIRS No. 2 small mammals, and plants within the inter- mountain region. Special reviews are pres- ented on Artemisia, Atriplex, and the genus Eriogonum and its relatives. Several broad biological problem areas within the region are reviewed, including the nature of the floristic transition zone between the Mojave and Great Basin deserts, the endemic flora of the Colorado Plateau, the distribution of the juniper-pinyon community in the Great Basin, and the evolutionary development of the alpine biota of the Intermountain Re- gion. Special considerations are given to the problems of managing native plant and ani- mal populations in the area. The general public will be especially in- terested in the role biogeographic consid- erations are now playing in the under- standing of the present-day distribution of organisms within the Intermountain Region and the management options available for the use and protection of these vital natural resources. Public land managers will find that many of the biogeographic principles discussed will help them to better under- stand the nature of arid land resources, en- dangered species, and the impact of man's technology in the American West. Each chapter stands as a unit with an in- troduction, a discussion, and a summary of pertinent literature. An index completes the volume. The Intermountain Biogeography Sym- posium was held at the University of Mon- tana, Missoula, 14-15 June 1976. The sym- posium was sponsored by Brigham Young University and the Intermountain Forest and Range Experiment Station of the U.S. Forest Service, with the cooperation of the Botanical Society of America and the Pacif- ic Section of the American Association for the Advancement of Sciences. The sym- posium was arranged by Kimball T. Harper of Brigham Young University and Ralph C. Holmgren of the U.S. Forest Service, with the assistance of Dr. George Edmunds of the University of Utah, Dr. Joseph Murphy of Brigham Young University, Dr. Neil West of Utah State University, and Edward Smith of the Bureau of Land Management. Major financial support for the symposium and the publication of these proceedings has been provided by Brigham Young Univer- sity and the U.S. Forest Service. We thank Dr. Stephen L. Wood, editor of the Mem- oirs series of the Great Basin Naturalist and the staff of Brigham Young University Press for their help in bringing the book to pub- lication. The cover was designed by Kaye H. Thorne. K. T. Harper and J. L. Reveal Symposium Organizers. THE BIOTA OF THE INTERMOUNTAIN REGION IN GEOHISTORICAL CONTEXT Arthur Cronquist' Abstract.— The present Great Basin Floristic Province had achieved roughly its present topographic con- formation by some time in the Miocene epoch and had a climate not too different from the present one, though probably a little warmer, moister, and less continental. Both the flora and the fauna took on a fairly modern as- pect during the Miocene, as a result of worldwide evolutionary changes and more specific adaptation to the con- ditions of the region. Changes in the biota since that time mainly reflect evolution and migration at the level of species and, to a lesser extent, genera, in response to regional conditions and the repeated fluctuations in climate. The climatic reversals of the Pleistocene caused repeated inverse migrations of more northern, mesophytic ele- ments in the flora, on the one hand, and more southern, xerophytic elements on the other. These expansions and contractions of range favored hybridization and genetic mixing among related plant species. The fauna of the re- gion, dependent eventually on the flora, must have been subjected to basically the same set of repeated changes in range and local distribution during the Pleistocene. About 10,000 years ago many of the large mammals in the Intermountain Region, as elsewhere in North America, rapidly became extinct, perhaps largely through overkill by primitive man. A proper understanding of the present is always facilitated by some knowledge of the past. Therefore I want to say something about the geological and biological history of the Intermountain Region, to help pro- vide a proper setting for the other papers of this symposium. Nearly everything that I have to say is already in the scientific liter- ature somewhere, but the particular syn- thesis may be in part new. As a first approximation of the truth, one may say that the aspect of the vegetation of any region is controlled by the climate, and the taxonomic composition of the flora is determined by the climate and the history. The general nature of the fauna is in turn determined by the vegetation, and the tax- onomic composition of the fauna is deter- mined by the vegetation and the history. It is, of course, also true that the fauna influences the flora. One of the reasons that grasses predominate in certain climates is that they are better adapted to withstand grazing than are most other herbaceous plants. Furthermore, evolution of floral structure is to some extent correlated with the evolution of pollinating insects (Leppik 1957), and particular species of plants may become dependent on particular pollinators. A notable example that may be familiar to many readers is provided by Yucca and the Tegeticula moth. Some of the importance of the influence of the fauna on the flora is also shown by the devastating effect of the introduction of goats to some of the islands off the Pacific Coast of southern North America. More complex interactions be- tween plants and animals also occur. Yet the preponderant control is that exerted by the food makers (plants) on the food eaters (animals). Therefore it is reasonably possible to consider the vegetation and flora of a re- gion with only secondary attention to the fauna, whereas any proper consideration of the fauna must be grounded in a knowledge of the vegetation. These facts, or what I take to be facts, are fortunate for me, be- cause I know a lot more about plants than I do about animals. The Intermountain Region may be vari- ously delimited. For purposes of this dis- cussion, I take its limits to be those of the The New York Botaima! Carilen. Bronx, New York 10458. GREAT BASIN NATURALIST MEMOIRS No. 2 Great Basin Floristic Province, as defined by Gleason and Cronquist (1964). In large part these limits are the same as those of the Intermountain Flora (Cronquist et al. 1972), but the Great Basin Floristic Pro- vince extends somewhat further south in- to Arizona and also includes a part of north- western New Mexico as well as a sliver of western Colorado. In addition to the hydro- graphic Great Basin, the area under consid- eration also includes the Snake River Plain and the more westerly segment of the Colo- rado Plateau. The region has a continental climate, with fairly hot, dry summers, and cold, snowy winters. The lowlands and foot- hills are largely desert and semidesert; a more mesophytic flora often occupies the upper elevations. South of the Inter- mountain Region lie hotter deserts, marked especially by milder winters. These southern deserts have a rather different flora, and the plant communities are often dominated by Larrea. The southern deserts are not a part of the Intermountain Region as here de- fined. Geologic History We shall start our consideration of the in- termountain biota with a summary of the geologic history of the region from the Cre- taceous period to the present. Much of the information in this section comes from pa- pers by Bateman (1968), Eardley (1968), and Roberts (1968). The present Intermountain Region has been at middle latitudes since before the beginning of the Cretaceous. North America has drifted westward, with respect to Eu- rope and Africa, throughout that time, but the latitude of our area has changed rela- tively little (Dietz and Holden 1970, Smith et al. 1973). Our region has been subjected to repeat- ed and almost continuous tectonic distur- bance, leading to uplift and erosion, from the beginning of the Cretaceous to the pres- ent. The terrain throughout that time has been highly varied, doubtless producing a diversity of habitats. The Upper Cretaceous, in particular, was a time of great and pro- longed uplift in western Utah and eastern Nevada. There was a large interior drainage basin in north-central Nevada even in the late Upper Cretaceous, and in mid-Eocene time there appear to have been high moun- tains and large lake basins throughout most of the present hydrographic Great Basin. In Oligocene time these lake basins were con- siderably elevated and themselves subjected to erosion. The present Rocky Mountains and Colo- ado Plateau began to rise early in the Ter- tiary, and they have continued to rise at varying rates until the present. The Sierra Nevada, bounding the Great Basin on the west, also has a long history. After a rela- tive quiescence during the Oligocene, the tilt-uplift of the Sierra Nevada was consid- erably accentuated during the Miocene. The concurrent uplift of the Rocky Mountains and Colorado Plateau shaped the Great Ba- sin. By some time in the Miocene, it ap- pears that "basin and range topography ex- tended from the Wasatch Mountains to the Sierra Nevada, and most of the area drained into interior basins" (Roberts 1968). There is some difference of opinion on the timing, however, and Axelrod (1950) believes that the present interior drainage of the Great Basin dates from near the close of the Pliocene. The Snake River Plain, forming a broad crescent across southern Idaho, belongs to the Great Basin floristic province but is geologically distinctive. The Upper Cre- taceous uplift in western Nevada and east- ern Utah extended across the present Snake River Plain as well. During Eocene time the present Snake River Plain was buried by lava in a major and prolonged tectonic- disturbance that formed a volcanic plateau extending from western Wyoming across southern Idaho and probably into eastern Oregon. In Oligocene time the Snake River basin took shape, possibly "as a tension rift in the lee of the Idaho batholith" (Axelrod 1968), which began to drift north. Sub- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM siclence of the basin and outpouring of new lava flows have continued until the present time. The youngest deposits in the Craters of the Moon region at the north edge of the Snake River Plain are probably only a few hundred years old. It is thought that for most of the Cre- taceous period the climate of the world was relatively warm and equable, and that trop- ical and subtropical climates entirely suit- able for the growth of forests extended from about 60 degrees north to about 60 degrees south (Barnard 1973). As late as the Eocene, the London Clay flora, at 35 degrees north, is definitely tropical (Hughes 1973). One may reasonably have some doubt about how humid the climate may have been in Ne- vada during the Upper Cretaceous, because of the presence of an interior drainage ba- sin, but such Cretaceous fossil floras as we have from the western United States suggest the presence of adequate moisture. The frequent presence of interior drain- age basins in the Intermountain Region for many millions of years past tells us some- thing about the climate. The precipi- tation/evaporation ratio for much of the re- gion much of the time must have been something less than 1. At a p/e ratio of more than 1, lake basins fill and spill over, finding external drainage. It is generally considered that a p/e ratio of not less than about 1 is required to support a forest. Therefore, for much or most of its span of existence the Great Basin is not likely to have been widely forested. Other parts of the Intermountain Region appear to have had a similar climatic regimen. Island Topography The topographic diversity of the Inter- mountain Region, with its associated differ- ences in temperature and moisture, effec- tively converts the habitats for many species of plants and animals into a series of is- lands. Not only the mountains, but also the valleys, form such islands for species with- out good means of dispersal. Birds can trav- el from one island to another, but small mammals frequently cannot. Different kinds of plants likewise differ in the ability to pass the inhospitable stretches between is- lands. On the other hand, these islands do not have the relative permanence of oceanic is- lands. The various island habitats in the In- termountain Region have expanded and merged, contracted and broken up, dis- appeared and reappeared, during Pleisto- cene and post-Pleistocene time because of changes in the climate. The principles of is- land biogeography, as expounded for ex- ample by Mac-Arthur and Wilson (1967), are pertinent to the Intermountain Region, but their effect is limited by the climatically controlled changes in island area. Early Angiosperm Evolution The angiosperms appear to have origi- nated early in the Cretaceous. Since we do not have fossils to connect the angiosperms to their necessarily gymnospermous ances- tors, we cannot say with certainty that they did not originate somewhat earlier. The fos- sil record does make it clear that the evolu- tionary diversification of the group did not get well started until the Cretaceous. Ang- iosperms enter the fairly early Lower Cre- taceous fossil record as an uncommon and not highly diversified group. Many of these early angiosperm fossils were at first opti- mistically identified with modern genera, leading to the widespread belief that the angiosperms entered the fossil record full- blown. We can now say with some assur- ance that the reverse is true. The pollen record speaks eloquently to the relative ho- mogeneity of the early angiosperms, and a reexamination of the megafossils shows that their identification with modern genera was disastrously incorrect. The purportedly Ju- rassic palm from Utah (Tidwell et al. 1970) is clearly a palm, but it is not Jurassic. The stratigraphy of the site where it was collect- ed is complex, and subsequent careful study shows that it is of Tertiary age (Scott et al. 1972). 6 GREAT BASIN NATURALIST MEMOIRS No. 2 The comments made in this paper on angiosperm evolution in general are heavily influenced by studies in the past decade by Dilcher (1969, 1973), Doyle (1969), Hickey (1973), Walker and Doyle (1975), Doyle and Hickey (1976), Hickey and Wolfe (1975), and Wolfe et al. (1975), who are in the forefront of the ongoing reevaluation of the early angiosperm fossil record. Their pub- lished work and my conversations with them have helped to shape my views, and I am particularly indebted to Dr. Leo Hickey for advice and counsel during the prepara- tion of this paper. Within the Inter- mountain Region, the work of Axelrod (1948, 1950, 1952, 1956, 1958, 1964, 1966, 1968, 1975) is of course preeminent. With- out it, our knowledge of the fossil flora would be scanty indeed. The interpretation presented is, as always, my own; those who helped me are not to be held responsible for what I might say. The place of origin of the angiosperms is still uncertain. It is clear that they are basi- cally a tropical group, but beyond that the situation is debatable. We can say that as early as the Aptian stage of the Lower Cre- taceous, 125 million or more years ago, they were well scattered in both Gon- dwanaland and Laurasia, including North America, but that they did not begin to dominate the landscape until the Upper Cretaceous. There is no reason to suppose that the Intermountain Region had anything to do with the origin of the angiosperms, but at the same time it is clear enough that it has supported some angiosperms at least from the Albian stage of the lower Cre- taceous to the present. Unfortunately we can not yet see a his- torical connection between the Cretaceous and Tertiary angiosperm floras of the Inter- mountain Region, or indeed of most other parts of the world. Most of the Cretaceous genera did not persist long if at all into the Tertiary, and the limited fossil record does not show whether our early Teritary genera originated in situ from the Cretaceous ones or migrated in from elsewhere. One of the- few Upper Cretaceous fossil floras definitely known from within the Intermountain Re- gion is in the Blackhawk formation in cen- tral Utah, a member of the Mesa Verde Group (Parker 1968, as reported by Tidwell et al. 1972). This flora included some palms and a number of woody dicotyledons and is thought to indicate humid lowland condi- tions under a warm-temperate to sub- tropical climate. This is in general harmony with views of the Cretaceous climate of the region based on other data (e.g. Axelrod 1950). Beginning with the Paleocene, we have a more nearly continuous history of the Inter- mountain flora, but even so there are some considerable gaps. Well over half of the Pa- leocene genera of angiosperms in the world flora are now extinct, and the fossil record as studied to date rarely shows the origin of modern genera from the more archaic ones. It appears that in the Paleocene the climate of the Intermountain Region was still reasonably warm and moist, sub- tropical or warm temperate, as it had been in the Upper Cretaceous. Evolution of Floristic Groups in the Intermountain Region By the middle of the Eocene, some 50 million years ago, the climate in the Inter- mountain Region had begun to dry out. The first indication of this in the fossil record comes in the Eocene Green River flora of northwestern Colorado and northeastern Utah (Axelrod 1950, MacGinitie 1969). This resembles the early Oligocene Florissant flora from Colorado (MacGinitie 1953) and like it may have been a subtropical sa- vanna-woodland. No closely similar flora ex- ists today. Drying of the Intermountain climate con- tinued, with some fluctuations, throughout the Tertiary. By early Oligocene the mean temperature of the world, at least in pres- ently temperate regions, had begun to drop (Bowen 1966), and in late Oligocene it dropped markedly (Wolfe and Hopkins 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 1967); it never again regained the Cre- taceous levels. Concomitant with increasing aridity and decreasing mean temperature in the Intermountain Region was a gradual trend toward a more continental climate, with hot, dry summers and cold, somewhat moister winters, continuing until about the middle of the Pliocene. Floristic changes in the Intermountain Region were, of course, related to the evo- lutionary diversification of the angiosperms throughout the world. The monocotyledons evidently diverged from primitive dicotyle- dons shortly after the appearance of angio- sperms in the fossil record, during the Lower Cretaceous. Palms became important elements of the world flora during the Cre- taceous, and grasses in the Oligocene or earlier. Dicotyledonous herbs were rare throughout the Cretaceous and on into the Paleocene and Eocene. They began to be- come more abundant in the Oligocene, and they increased dramatically at the beginning of the Miocene, some 25 million years ago. During Miocene time the flora of the world began to take on a fairly modern aspect, with a great many genera that still exist today. The increase in dicotyledonous herbs in the mid-Tertiary is thought to reflect at least in part the increasing aridity of the climate throughout much of the world, en- larging the area not suitable for forests. It is evident that during the drying of the climate in western North America the Ter- tiary flora sorted itself out into a more northern, mesic flora dominated by trees, and a more southern, xeric flora with few if any trees. Fossil floras from near the Oligo- cene-Miocene boundary in southwestern Montana suggest a mainly forest vegetation, with some elements from the drylands to the south (Becker 1969). Within the dryland flora there was a further differentiation into a more northern segment adapted to cold winters, and a more southern segment adapted to a warmer climate. The present Great Basin Floristic Province, representing the more northern of these two dryland floras, evidently took shape in the Miocene. Indeed the Miocene boundary between the Great Basin flora and the Mohave Desert (a part of the more southern flora) may have been about where it is now (Axelrod 1950). It is not clear how much of the differen- tiation of the intermountain flora during the Tertiary represents evolution in situ, and how much of it reflects immigration from other regions. Certainly both processes oc- curred. A similar sorting out occurred in other parts of the world, and in Eurasia this involved many of the same families and even genera. It is not likely that the same taxonomic groups originated independently in North America and Asia. There must have been some interchange. The genus Artemisia might be considered in this regard. Although Artemisia tridentata Nutt. and its immediate allies dominate the scene in much of the Intermountain Region, Artemisia is not of American origin. The tribe Anthemideae of the family Asteraceae (Compositae), to which Artemisia belongs, is basically an Old-World tribe, and most of the species of Artemisia itself occur in the Old World rather than in the new. Arte- misia and Juniperus characterize the land- scape in parts of Armenia, for example, as well as in the Intermountain Region. Arte- misia in the western United States is an im- migrant, although the particular species we now have may well have originated here from immigrant ancestors. Some other members of the Asteraceae are definitely American. The whole tribe Heliantheae is clearly so. Its present center of diversity is in the arid highlands of cen- tral Mexico, and it seems reasonable to sup- pose that the tribe is of Mexican or western American dryland origin. Many members of the group here will probably be acquainted with species of Balsamorhiza, Chaenactis, Enceliopsis, Eriophyllum, Viguiera, and Wyethia, all members of the Heliantheae, that grow in the Intermountain Region. The large genus Haplopappus, in the tribe As- tereae, is strictly American (North and South), with one center of diversity in west- ern North America and another in 8 GREAT BASIN NATURALIST MEMOIRS No. 2 Chile. Erigeron is another large genus of the Astereae that has its principal center of di- versity in western North America and ap- pears to have originated there. I am not suggesting that these several genera of He- liantheae and Astereae originated in the In- termovmtain Region, but they probably did not have far to come to get here. Atriplex, another important genus in the Intermountain Region, has more species in the Old World than in the New. The family Chenopodiaceae, to which Atriplex belongs, has considerable concentrations of species in the Mediterranean region, in western and central Asia, in South Africa, and in Austra- lia, as well as in the drier parts of both North and South America. The Roraginaceae appear to be tropical and woody in origin, but they are well rep- resented by numerous herbaceous genera and species not only in our arid West but also in the Mediterranean region and in central Asia. To what extent did our boragi- naceous intermountain herbs originate in North America from tropical woody ances- tors, and to what extent do they reflect im- migration of herbs from the Old World? Certain genera, such as Cryptantha and Plagiobothrys, are clearly American now, whatever their eventual origin, but others, such as Lithospermum, are well developed in Eurasia and may well be immigrants in North America. The Rrassicaceae are well represented in the Intermountain Flora, but they are even more numerous and diversified in the arid region from central Asia to the Mediterra- nean, and the family as a whole is probably of Old World origin. Such familiar genera as Lesquerella, Physaria, Stanleya, Strep- tanthus, and Thelypodium are strictly American, whatever the origin of the family as a whole. Cardamine, Lepidium, and Ror- ripa, on the other hand, are well represent- ed in the Old World also. A few families, such as the Hydro- phyllaceae and Polemoniaceae, evidently have their principal center of diversification in western North America, even if the re- gion of their ultimate origin is not yet clearly established. Such genera as Phacelia, in the Hydrophyllaceae, and Cilia, in the Polemoniaceae (Grant 1959), are clearly at home in the Intermountain Region. There is no reason to suppose that they came in from some other continent. Axelrod and Chaney have in various pa- pers (e.g., Axelrod 1958) promoted the thought that the Tertiary flora of the west- ern United States can be divided into an Arcto-Tertiary and a Madro-Tertiary seg- ment. The Arcto-Tertiary geoflora, domi- nated by deciduous trees, is considered to have been very wide-spread, extending across most of northern North America and northern Eurasia. The deciduous forest of the eastern United States is considered to be the nearest modern American counterpart and a lineal descendant of the Arcto-Ter- tiary geoflora. The Madro-Tertiary geoflora, on the other hand, was adapted to drier, warmer conditions, with many xeromorphic shrubs, the trees being restricted to favor- able habitats, or completely wanting. The Madro-Tertiary geoflora as so conceived was geographically more restricted than the Arcto-Tertiary, being confined to northern Mexico and the southwestern United States. The "Madro" part of the name comes from the Sierra Madre Occidental in north- western Mexico. The Madro-Tertiary flora is considered to have originated in situ from subtropical western American plants that gradually became adapted (through evolu- tion) to xeric conditions. The concept of Arcto-Tertiary and Madro-Tertiary floras has recently been challenged by a number of authors, notably Wolfe (e.g., 1969), and is now in some dis- repute. The problem, to my mind, is that some useful generalizations have been taken too literally and interpreted too rigidly. I am reminded of Gleason's challenge (1926) to Clementsian concepts of plant associ- ations. Most modern ecologists agree with Gleason that the association is a mental construct that can be defined only arbi- trarily. The idea that the community is an 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 9 organism is a good aphorism, but it can lead to serious misunderstanding if it is taken literally. Likewise the concept of an Arcto-Tertiary and a Madro-Tertiary geof- lora is useful if one conceives of these floras broadly and loosely and recognizes that each of them encompasses a considerable amount of diversity, that some elements were common to both, and that there was continuous interchange between them. It is helpful to think in terms of floristic groups, but we should keep it constantly in mind that each species has its own limits of eco- logical tolerance, its own means of migra- tion, and its own evolutionary potentialities, the last being influenced also by hybridiza- tion with related species. The species that make up any floristic group have entered that group, through immigration or through evolution in situ, at various times in the past, and species that are now associated may not' remain associated under some fu- ture climatic regimen. I can easily agree with Axelrod that the modern desert flora of the western United States and northern Mexico probably "de- veloped during the Tertiary period by grad- ual adaptation of more mesic plants to slowly expanding dry climate" (Axelrod 1950). It seems perfectly logical to suppose that the present flora of the warm deserts south of the Intermountain Region is a line- al descendant of a Madro-Tertiary geoflora that differentiated originally from American plants adapted to similarly warm but more mesic climates. There is no other likely source. Some of them doubtless originated instead by adaptation of Arcto-Tertiary taxa to warmer, drier climates, and some of these that entered the warm deserts from the north doubtless take their origin eventu- ally in Asia, but it strains credulity to de- rive the bulk of the Madro-Tertiary flora in such a way. The origin of the Arcto-Tertiary flora is a more difficult question. Obviously it repre- sents an adaptation of tropical or sub- tropical plants to a cooler but still moist climate. Since it extended across both North America and Eurasia, one cannot a priori assume that it came principally from either an Old- World or a New World source. It seems logical to suppose that the Arcto-Ter- tiary flora originated from the Cretaceous tropical and subtropical Laurasian flora, but at the present time that is pure speculation. We have noted that the angiosperm fossil record as presently understood does not provide a good connection between the Cretaceous and the Tertiary. The problem is complicated by the fact that during the Mesozoic era and most of the Tertiary peri- od North and South America appear to have been separated, not contiguous. South America was part of the southern continent, Gondwanaland, whereas North America was part of the northern continent, Laurasia. North America drifted away from Europe during and after the Cretaceous, but, until recently, it has mostly been well separated from South America. I say "mostly," be- cause the geologic history of the Caribbean is complex and insufficiently understood, and the possibility of a direct connection between North and South America at some time during the Cretaceous or early Ter- tiary cannot be completely discounted. At the present time the tropical part of the flora of North America (as represented by southern Florida, the West Indies, south- ern Mexico, and Central America) is clearly allied to the flora of South America. If there is any surviving Laurasian element in the present tropical North American flora, it is so thoroughly amalgamated into the Gondwanaland, South American flora that no one has yet been able to recognize it. In making this statement I exclude from con- sideration some primarily temperate-zone species and genera that extend into the tropics at the southern limit of their range. Although the vegetation of most of the Intermountain Region is rather similar in as- pect to that of the deserts farther south, it is very different in floristic composition. As Axelrod (1950) has pointed out, some of the dominant genera in the Intermountain Re- gion, such as Artemisia, Atriplex, and Cera- 10 GREAT BASIN NATURALIST MEMOIRS No. 2 toides (Eurotia), apparently relate to the Ar- cto-Tertiary rather than the Madro-Tertiary flora. Likewise Astragalus, one of our larg- est genera in terms of number of species, has an even larger number of species in dryland Eurasia. Even if one prefers to avoid the terms Arcto-Tertiary and Madro- Tertiary, these genera still relate to Asian desert plants, presumably by way of a Ber- ingian connection, rather than to plants from farther south in western North Ameri- ca. On the other hand, such large genera as Penstemon and Eriogonwn are strictly American, best developed in arid western North America, without any obvious in- dication of a more southern (Madro-Ter- tiary) origin. Haplopappus may well be from the Madro-Tertiary, as Axelrod sug- gests, but its derivative Chrysothamnus cen- ters in the Great Basin. We have already noted that some of the common genera of Heliantheae in the Intermountain Begion may well be of Madro-Tertiary affinity. Thus it is not possible to assign the charac- teristic flora of the Great Basin province to either a chiefly Madro-Tertiary or a chiefly Arcto-Tertiary origin. Both of these Tertiary floras clearly contributed to the present flora of the region. Thus, by some combination of differen- tiation from native elements, immigration from near and far, and proliferation of the immigrants, the Intermountain Flora ac- quired its special character during the Miocene epoch. Xerophytes predominated especially at lower elevations, but meso- phytes survived in the moister habitats, of- ten at higher elevations. These two types have been in continuous competiton in the Intermountain Begion since that time. Tension between Mesophytic and Xerophytic Communities Although the Great Basin floristic province took shape in the Miocene, it was not immediately so dry as it is now. Axelrod (1948) considers that open environments ex- tended through the region in the Middle Pliocene, but that the plant community was predominantly grassland, with semidesert shrubs on the drier slopes. In Miocene and Pliocene time the presently desert regions supported species comparable to those in the pinyon-juniper woodland and oak wood- land that now occur at slightly higher ele- vations or around the borders of the desert. Axelrod (1950) considers that the trend toward a drier, more continental climate in the Intermountain Begion, begun early in the Tertiary, culminated in Middle Pliocene time, perhaps 4 or 5 million years ago. Lat- er in the Pliocene the climate probably be- came a bit cooler and moister. The Pleisto- cene, as we all know, was marked by alternating glacial and interglacial stages. From a long-term geohistorical viewpoint, the present time may be merely another Pleistocene interglacial. Actual glaciers in the Intermountain Begion were largely re- stricted to upper elevations in the moun- tains; the continental ice sheet did not reach that far south in western North America. The glacial periods were times of rela- tively lower temperatures and higher p/e ratio in the Intermountain Begion, cooler and more mesic than the interglacials. Dur- ing the glacial periods, the mesophytes, many of them of northern floristic affinities, expanded their distribution at the expense of the xerophytes; in the interglacials the process was reversed. The great differences in elevation, together with the strong local differences in moisture relations according to slope and edaphic factors, combined with the repeated shifts in climate to keep the species populations in constant turmoil throughout the Pleistocene. T. M. Barkley (personal communication) has suggested that the blurred boundary between Senecio strep- tanthifolius (a highland species) and Senecio multilobatus (a lowland, more xerophytic species) in Utah reflects such hybridization. Local polyploidy helps such hybrids and hybrid segregates to persist in appropriate habitats. Another example of the advance and re- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 11 treat of species in the Intermountain Region due to climatic changes is provided by the oaks. In north-central Utah there exist today clones of oak that have been conclusively demonstrated to be hybrids between Quercus gambelii and O. turbinella. Quercus gamhelii is common in the area today, but Q. turbinella reaches its present northern limits more than 250 miles to the south of these hybrids. It is reasonably believed that, during the postglacial hypsithermal period, some five or six thousand years ago, the range of Q. turbinella extended north into north-central Utah, permitting the forma- tion of the hybrids (Cottam et al. 1959). Alpine fir, Abies lasiocarpa, provides an example in the reverse direction. According to Cottam et al. (1959), fossils discovered in 1957 by D. J. Jones demonstrate that alpine fir grew along the shores of Lake Bonne- ville at a time when the lake level stood well below the Provo stage. Recent fluctua- tions in the level of Great Salt Lake remind us that p/e ratios in the Intermountain Re- gion continue to fluctuate, but up until now the climatic changes during the relatively short time for which we have formal, writ- ten records do not approach the magnitude of the changes that occurred during geolog- ic time. Present-Day Correlation of Elevation with Floristic Groups Elevation is closely correlated with mois- ture relations as well as with temperature in the Intermountain Region. As one goes higher into the mountains, the temperature drops and the p/e ratio increases, and one finds a progressively more northern element in the flora. Many years ago I read some- where that in the western United States one can roughly equate one mile of latitude with four feet of altitude. In my own expe- rience, this conversion factor works fairly well, although there are, of course, always modifying factors to be taken into account. At moderately high elevations in the moun- tains, one finds many species similar or identical to those of the northern coniferous forest, and above timberline one finds many species similar or identical to those of the modern circumboreal arctic flora. The spruce-fir forests of midupper elevations in the Intermountain Region represent a south- ern extension of the northern coniferous for- est. Even though the dominant species are different, they compare closely with species from the northern forest. Abies lasiocarpa compares with Abies balsmnea, Picea engel- mannii and P. pungens compare with P. glauca, and Pinus contorta compares with P. banksiana. Pseudotsuga menziesii, on the other hand, does not have a boreal equiva- lent. North of the Intermountain Region, in the northern Rocky Mountains of Canada and the northwestern United States, a very large proportion of the high-mountain spe- cies can be related directly to something from the holarctic or the northern con- iferous forest. As one goes progressively southward, a larger and larger proportion of the alpine species are evidently highland derivatives from common lowland elements. In the Intermountain Region both of these types are well represented at upper alti- tudes. Alpine and subalpine species of Are- naria, Gentiana, Myosotis, Pedicularis, Ranunculus, and Saxifraga are likely to have boreal affinities. On the other hand, montane species of Allium, Eriogonum, Hul- sea, Hymenoxys, Lomatium, and Penstemon, even at the highest elevations, generally re- late to species of lower elevations, often of dry habitats. Some common montane gen- era, such as Erigeron, do not fit into either of these patterns. Erigeron is best developed in the western American cordillera, but the species of the dry lowlands are evidently advanced, and the more primitive species are distinctly mesophytic. Evolutionary History of Mammals The evolutionary history of the mammals parallels in many ways that of flowering plants. Although the group takes its origin 12 GREAT BASIN NATURALIST MEMOIRS No. 2 from therapsid reptiles in the Triassic peri- od, the placental mammals do not enter the fossil record until late in the Cretaceous. Placental mammals diversified explosively during the Paleocene, and they have been the dominant animals in terrestrial ecosys- tems since that time. Evolution of mammals in North America is closely correlated with that in Eurasia, but not well correlated with that in South America, because of the essen- tial separation of North and South America until relatively recent times. The animals that may have had the most important influence on the plants during the Tertiary period were the grazing ani- mals—ungulates, in the broad sense. Grazing mammals began to evolve in the Paleocene or Eocene, and they reached full flower in the Oligocene and Miocene (Jones and Arm- strong 1973). One may reasonably suppose that there is a relationship between the evo- lution of grazing mammals and the rise of grasses during the same general time. Grasses originated no later than the Oligo- cene, and by Miocene time they were com- mon. The intercalary meristem of the grass leaf can reasonably be interpreted as an adaptation to grazing pressure. Thus, al- though it may be true that at any given time the nature of the fauna is more depen- dent on the flora than vice versa, in the long rim the evolution of plants is strongly influenced by animals. The most startling feature of the evolu- tionary history of mammals in North Ameri- ca was the rapid extinction of a great many of the large mammals about ten thousand years ago. There is no real parallel in the evolutionary history of plants. Similar ex- tinction occurred to varying degrees in oth- er parts of the world, least of all in Africa. Both climatic changes and the influence of early man have been invoked to explain the massive extinctions. In North America, the case for the predominant influence of man is very good (Martin 1967), although the subject still evokes considerable debate and difference of opinion (Axelrod 1967). The large mammals had survived much more ex- tensive climatic changes during the Pleisto- cene, and their disappearance from the scene appears to be closely correlated with the spread of man. Human hunters killed the large herbivores, and many of the large predators disappeared along with their prey. Bison, camels, elephants, and horses were abundant in the Intermountain Region dur- ing the Pliocene and Pleistocene (Axelrod 1950), but of these only the bison survived the human onslaught. It is clear enough that horses, at least, are well adapted to modern conditions in the Intermountain Region, and burros do very well a little farther south. On the other hand, the camels introduced into our southwest more than a century ago did not make the grade, although they might well have done so in the absence of Evolutionary History of Birds The birds apparently originated in the upper Jurassic and began to radiate in the Cretaceous, but nearly all the Cretaceous families are now extinct. After the extinc- tion of the dinosaurs and before the evolu- tion of large carnivorous mammals, there were some large flightless birds, which played the ecological role later taken over by large mammalian carnivores. In the northern hemisphere these birds were com- mon from the Upper Paleocene to the middle of the Eocene. An ecologically sim- ilar but taxonomically distinct group of large, flightless, predatory birds was com- mon from early Eocene to middle Pliocene time in South America, an area into which the large carnivores made a relatively late entry. Again the geographic separation of North America from South America during most of the time from the Cretaceous until late in the Tertiary had a profound effect on evolutionary patterns. By the end of the Eocene the birds were highly diversified, and all living families and orders can be traced back at least that far. In Miocene time the avifauna began to take on a more modern aspect, and most of the 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 13 modern genera had come into existence by Pliocene time (Storer 1974). There was no great wave of recent extinction comparable to that of the large mammals. The avifauna of the Intermountain Re- gion has no endemic species and is distin- guished mainly by what isn't there. The species are all more or less widespread. Dis- tinctively Californian, southern Rocky Mountain, and Mohavean species mostly do not extend into the Intermountain Region (W. H. Behle, this symposium). Evolutionary History of Insects The evolutionary radiation of insects, like that of flowering plants, mammals, and birds, goes back many millions of years. The Coleoptera (beetles) are well known as fos- sils as far back as the Permian period. The Diptera and Hymenoptera date from the Jurassic period, but only in forms such as midges and crane flies (Diptera) and saw flies (Hymenoptera), which are not and pre- sumably never were important pollinators. The bees and the higher Diptera, which are now important pollinators, first appear in the fossil record in early Tertiary time, al- though they may well have originated somewhat earlier (Carpenter 1953, Baker and Hurd 1968). Most or all of the early Tertiary bees belong to extinct genera, and one may legitimately speculate that the ev- olution of modern bees was intimately re- lated to the evolution of structurally com- plex, bee-pollinated flowers during the Tertiary. The Lepidoptera originated no lat- er than the late Cretaceous (MacKay 1970) and had already diversified to some extent in early Tertiary time, but here again the important pollinators are apparently not an- cient types. Drawing upon the recent dis- coveries of Cretaceous fossil insects report- ed by Rodendorf and Zherikhin in 1974, Doyle (1976) visualizes "major extinctions of 'Jurassic' groups within a relatively brief in- terval of the Late Cretaceous, and a some- what slower rise of groups now associated with angiosperms." The coevolution of structurally complex flowers and insects ca- pable of recognizing complex patterns rep- resents another example of major evolution- ary interaction between plants and animals (Leppik 1957, Baker and Hurd 1968). The faunistic differentiation between Laurasia and Gondwanaland shows up at least in the aquatic insects of the Inter- mountain Region. Species of Gondwanaland ancestry occur mostly in the warmer wa- ters, or their eggs hatch relatively late in the summer. Some species of Laurasian af- finity connect to Eurasia through Beringia, and others through Europe (G. F. Edmunds, personal communication). Literature Cited Axelrod, D. I. 1968. Climate and evolution in western North America during Middle Pliocene time. Evolution 2: 127-144. 1950. Evolution of desert vegetation in west- em North America. Carnegie Inst. Washington Publ. 590: 215-306. 1952. A theory of angiosperm evolution. Ev- olution 6: 29-60. 1956. Mid-Pliocene flora from west-central Nevada. Univ. Calif. Publ. Geol. Sci. 33: 1-321. 1958. Evolution of the Madro-Tertiary geof- lora. Bot. Rev. 24: 433-509. 1964. The Miocene Trapper Creek flora of southern Idaho. Univ. Calif. Publ. Geol. Sci. 51: 1-148. 1966. The Eocene Copper Basin flora of northeastern Nevada. Univ. Calif. Publ. Geol. Sci. 59: 1-125. 1967. Quaternary extinctions of large mam- mals. Univ. Calif. Publ. Geol. Sci. 74: 1-42. 1968. Tertiary floras and topographic history of the Snake River Basin, Idaho. Geol. Soc. Amer. Bull. 79: 713-734. 1975. Evolution and biogeography of Madr- ean-Tethyan sclerophyll vegetations. Ann. Mis- souri Bot. Gard. 62: 280-334. Baker, H. G., and P. D. Hurd, Jr. 1968. Intrafloral ecology. Ann. Rev. Entom. 13: 385-414. Barnard, P. D. W. 1973. Mesozoic floras. In: N. F. Hughes (ed.), Organisms and continents through time. Publ. Syst. Assoc. 9: 175-187. Bateman, P. C. 1968 Geologic structure and history of the Sierra Nevada. Univ. Missouri Rolla J. 1: 121-131. Becker, H. F. 1969. Fossil plants of the Tertiary Beaverhead basins in southwestern Montana. Palaeontographica 127B: 1-142. 14 GREAT BASIN NATURALIST MEMOIRS No. 2 Billings, W. D. 1974. Adaptations and origins of alpine plants. Arctic Alp. Res. 6: 129-142. Bo wen, R. 1966. Paleotemperature analysis. Else- vier. Amsterdam. Carpenter, F. M. 1953, The evolution of insects. Amer. Sci. 41: 256-270. Cottam, W. P., J. M. Tucker, and R. Drobnick. 1959. Some clues to Great Basin postpluvial climates provided by oak distribu- tion. Ecology 40: 361-377. Cronquist, A., A. H. Holmgren, N. H. Holmcren, and J. L. Reveal. 1972. Inter- mountain Flora. Vol. 1. Hafner Publ. Co., New York. Dietz, R. S., and J. C. Holden. 1970. The breakup of Pangaea. Sci. Amer. 223(4): 30-41. Dilcher, D. L. 1974. Approaches to the identi- fication of angiosperm leaf remains. Bot. Rev. 40: 1-157. Dorf, E. 1969. Paleobotanical evidence of Mesozoic and Cenozoic climatic changes. Proc. N. Amer. Paleont. Conv. Sept. 1969 D: 323-346. Doyle, J. A. 1969. Cretaceous angiosperm pollen of the Atlantic Coastal Plain and its evolution- ary significance. J. Arnold Arb. 50:1-35. 1973. Fossil evidence on early evolution of the monocotyledons. Quart. Rev. Biol. 48: 399-413. 1976. Man bites botanical dogma. Paleobiol- ogy 2: 265-271. Doyle, J. A. and L. J. Hickey. 1976. Pollen and leaves from the mid-Cretaceous Potomac group and their bearing on early ang- iosperm evolution, pp. 139-206. In: C. B. Beck (ed.), Origin and early evolution of angiosperms. Columbia University Press, New York. Eardley, A. J. 1968. Major structures of the Rocky Mountains of Colorado and Utah. Univ. Mis- souri Rolla J. 1: 79-99. Gleason, H. A. 1926. The individualistic concept of the plant association. Bull. Torrey Bot. Club. 53: 7-26 Gleason, H. A., and A. Cronquist. 1964. The natural geography of plants. Columbia Univer- sity Press, New York. Grant, V. 1959. Natural history of the Phlox fam- ily. Martinus Nijhoff. The Hague, Netherlands. Hickey, L. J. 1973. Classification of the architecture of dicotyledonous leaves. Amer. J. Bot. 60: 17-33. Hickey, L., J, and J. A. Wolfe. 1975. The bases of angiosperm phylogeny: vegetative orphology. Ann. Missouri Bot. Gard. 62: 538-589. 62: 538-589. Hughes, N. F. 1973. Mesozoic and Tertiary distri- butions, and problems of land plant evolution. In: N. F. Hughes (ed.), Organisms and conti- nents through time. Publ. Syst. Assoc. 9: 188- 198. Jones, J. K., Jr., and D. M. Armstrong. 1974. Mammalia; pp. 401-416 in Encycl. Brit- annica, ed. 15, vol. 11:401-416. Leppik, E. E. 1957. Evolutionary relationship be- tween entomophilous plants and anthophilous insects. Evolution 11: 466-481. MacArthur, R. H., and E. O. Wilson. 1967. The theory of island biogeography. Princeton Univ. Press. MacGinitie, H. D. 1953. Fossil plants of the Floris- sant beds, Colorado. Carnegie Inst. Wash. Publ. 465: 83-151. 1969. The Eocene Green River flora of northwestern Colorado and northeastern Utah. Univ. Calif. Publ. Geol. Sci. 83: 1-203. MacKay, M. K. 1970. Lepidoptera in Cretaceous amber. Science 167: 379-380. Martin, P. C. 1967. Prehistoric overkill. In: P. S. Martin and H. E. Wright, Jr. (eds.), Pleistocene extinctions: the search for a cause. Proc. VII Congr. Internat. Assoc. Quaternary Res. 6: 75- 120. Raven, P. H., and D. I. Axelrod. 1974. Angio- sperm biogeography and past continental move- ments. Ann. Missouri Bot. Gard. 61: 539-673. Roberts, R. J. 1968. Tectonic framework of the Great Basin. Univ. Missouri Rolla J. 1: 101-119. RODENDORF, B. B., AND V. V. ZhERIKHIN. 1974. Paleontoogiya i okhrana prirody. Priroda 1974(5): 82-91. Scott, R. A. P. L., Williams, L. C. Craig, E. S. Barghoorn, L. J. Hickey, and H. D. MacGinitie. 1972. "Pre-Cretaceous" an- giosperms from Utah: evidence for Tertiary age of the palm woods and roots. Amer. J. Bot. 59: 886-896. Smith, A. G., J. C. Briden, and G. E. Drewry. 1973. Phanerozoic world maps. In: N. F. Hughes (ed.), Organisms and continents through time. Publ. Syst. Assoc. 9: 1-42. Storer, R. R. 1974. Bird. Encycl. Britannica, ed. 15, vol. 2: 1053-1062. TlDWELL, W. D., S. R. RUSHFORTH, J. L. REVEAL, AND H. Behunin. 1970. Palmoxylon simperi and Palmoxylon pristina: Two pre-Cretaceous angiosperms from Utah. Science 168: 835-840. TlDWELL, W. D., S. R. RUSHFORTH, AND D. Simper. 1972. Evolution of floras in the In- termountain Region, pp. 19-39. In: A. Cronquist, A. H. Holmgren, N. H. Holmgren and J. L. Reveal, Intermountain Flora, vol. 1, Hafner Publ. Co., New York. Walker, J. W., and J. A. Doyle. 1975. The bases of angiosperm phylogeny: Pa- 1978 INTERMOUNTAIN BIOCEOGRAPHY: A SYMPOSIUM 15 lynology. Ann. Missouri Bot. Gard. 62: 664-723. geny: Paleobotany. Ann. Missouri Bot. Gard. 62: Wolfe, J. A. 1969. Neogene floristic and vegeta- 801-824. tional history of the Pacific Northwest. Wolfe, J. A., and D. M. Hopkins. Madrono 2: 83-110. 1967. Climatic changes recorded by Tertiary Wolfe, J. A., J. A. Doyle, and V. M. land floras in northwestern North America. Page. 1975. The bases of angiosperm phylo- Symp. Pacific Sci. Contr. 25: 67-76. BIOGEOGRAPHY OF INTERMOUNTAIN FISHES Gerald R. Smith1 Abstract.— Eighty-three species of fishes belonging to 26 genera live in the area bounded by the Sierra Ne- vada, Grand Canyon, Rocky Mountains, and Snake River Plain. The waters inhabited by these fishes are part of the Great Rasin, Colorado River, Snake River, upper Pit River and upper Klamath River drainages. The adapta- tions and distribution patterns of these fishes have been shaped by extensional faulting and volcanic activity in the Great Basin, uplift of the surrounding ranges and plateaus, and cyclic fluctuations of the Pleistocene pluvials and interpluvials. Fossil evidence indicates that in the Pliocene many of the lineages had established distributions broadly inclusive of the present-day patterns, and the subsequent trends have been extinction and some species differentiation. Analysis of the fauna is based on designation of 48 barrier-bounded, faunally homogeneous drainage units and quantitative evaluation of patterns among species distributions and faunal similarities of drainages. Cluster analy- sis of species based on correlations among their patterns revealed the existence of a basic northern intermountain fluvial fauna consisting of Cottus bairdi, Prosopium williamsoni, Catostomus platyrhynchus, Rhinichthys cata- ractae, Richardsonhis balteatus, and their vicariants. These fishes have similar ecology and dispersal patterns. They are ecologically associated with Salmo clarki (upstream) and Rhinichthys osculus (downstream), but these two species have broader distributions, probably because of more frequent colonization via stream capture by the former and extinction resistance by the latter. Rhinichthys osculus is the most widespread intermountain fish, being found in 32 of the 48 drainage units; Gila bicolor is next most widespread, being found in 21 units. Fifty- one of the 83 species are found in only one drainage unit; 18 of these are endemic to that unit (33 are more widespread outside the study area). Species distributions are broader in the north, and northern and peripheral units have more species. The spe- cies:area curve for the Great Basin shows a steep slope (z = .59), with especially low species density in small drainages, indicating high extinction and low colonization. Postpluvial aridity, especially in the south, is the ma- jor cause of extinction and a major cause of isolation. Principal components and cluster analysis of drainage units, based on shared species, show a high correspondence between faunal similarity and geographic proximity (and weakness of barriers) and also reveal the effects of extinction in erasure of patterns. The Wasatch Range, Sierra Nevada, and southern divide boundary of the Snake River Plain have been strong barriers, leading to intensive faunal differentiation. Strong barriers and concomitant differentiation also exist in eastern Nevada, near the origi- nal center of Basin and Range tectonism. Two dozen examples of vicariant species are found to be associated with stronger-than-average barriers. The colonization rate and extinction rate have both been accelerated in postsettlement time by introductions and habitat destruction. Most drainages and populations have been affected. Five species and many local popu- lations have become extinct. Eighteen species and many more populations are vulnerable or threatened. Reversal of the trend will require sound ecosystem management of watersheds and restriction of exotic introductions. The distribution patterns of fishes in the survival has been dependent upon continu- intennountain region of the western United ity of lakes, marshes, springs, and streams in States offer an unusual opportunity to study isolated basins for many thousands of years, the evolutionary results of long-term ecolog- Dispersal has been dependent on occasional ical changes because of the insular nature of continuity of suitable habitat among basins, the drainage basins, the relatively well- Compared to most kinds of organisms, the known geological history, and the close de- restrictions on freshwater fishes are more pendency of these fishes on the continuity rigid; compared to most habitats, Great Ba- of their habitat in time and space. Their sin aquatic habitats have been less stable 'Museum of Zoology and Museum of Paleontology, University of Michigan, Ann Arbor, Michigan 48106. 17 L8 GREAT BASIN NATURALIST MEMOIRS No. 2 and more confined. These restrictions pro- vide relative constancy of some variables of interest and some replication in a quasi ex- periment of evolutionary responses to changing ecological conditions. The predominant environmental factors are the degree of isolation of the drainage basins and the late Cenozoic history of fluc- tuating pluvial and interpluvial episodes. The isolation has greatly restricted dispersal and colonization; the fluctuating climate has subjected the fish populations to extreme conditions, forcing adaptation or extinction. Fortuitously, the dominant aquatic habitats are also agents dominant in geomorpho- logical processes, and the lakes and streams that supported the fish populations have left stark and beautiful records in their wake (Gilbert 1890, Russell 1885). This paper is a summary of the relationship of fish distribu- tions to geological history and the contribu- tion of fish distributional data to the under- standing of that history, as reconstructed by Cope (1883), Snyder (1908, 1917), Hubb's and Miller (1948a,b), Miller and Hubbs (1960), Miller (1945, 1948, 1958, 1965), Hubbs, Miller, and Hubbs (1974) Evermann (1897), and La Rivers (1962). Geological Setting The Great Basin includes more than 150 drainage basins among approximately 160 regularly spaced, roughly parallel mountain ranges. The ranges and valleys trend north or northeast and are bounded by steeply dipping normal faults. Faulting began as early as Eocene or Oligocene, but most of the ranges were formed during the past 20 million years by crustal extension. The Great Basin crust is relatively thin and is bounded to the east, north, and west by zones of much thicker crust. Extension has been generally northwest-southeast and has been estimated between 50 and 300 km. Geophysical evidence indicates that the driving force may be a rising, spreading, semimolten body (diapir), whose origin is related to the early and Middle Cenozoic subduction of the Farallon plate at the West Coast trench (see Scholz et al. 1971, Atwater 1970). The structure and processes of the system are essentially those of an in- terarc basin. Volcanic evidence for this in- terpretation (from Armstrong et al. 1969) involves an episode of andesitic volcanism with much silicic ash in east-central Nevada 40-30 million years ago, abruptly changing to basaltic volcanism which radiated toward the margins of the Great Basin during the past 20 million years. Scholz et al. suggest that the outward radiation of basaltic vol- canism tracked the spreading margins of the underlying diapir and its concomitant crus- tal extension, the whole process being in- itiated at the release of compressional stress when the subduction of the Farallon plate was complete. The late Cenozoic tectonic relationship between the Great Basin and surrounding areas has been deduced from seismic data by Smith and Sbar (1974). An arcuate pat- tern of zones of shallow earthquakes defines the Great Basin boundary along the Snake River plain on the north, the Wasatch front on the east, and from about Cedar City west 200 km across southern Nevada on the south. Fault plane solutions indicate a slight counterclockwise rotation of the Great Ba- sin subplate accompanied by rifting along the Snake River Plain. Volcanic activity has proceeded eastward along the Snake River Plain at a rate of about 4 cm per year, the current focus being in the Yellowstone re- gion (Armstrong et al. 1975). This activity is interpreted by Eaton et al. (1975) and Smith and Sbar (1974) as marking the prog- ress of a mantle plume of molten magma being overridden by the westward-moving North American plate. This brief tectonic outline may be roughly correct, at least as far as the emp- irical surface history is concerned, and pro- vides us with an indication of several of the geological variables that probably played a part in the development of patterns of dis- tribution. It is not our intent to fit the dis- tributions to this history, but to recognize 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 19 the antiquity of the general topographic- pattern, the large potential for instability of drainage on a local scale, and the pervasive part played by volcanism. The evolution of structural features of Great Basin and Wasatch Range topography set the stage for increasing isolation of drainage units. The degree of isolation of basins has also increased with generally in- creasing aridity (Axelrod 1950). During Pleistocene pluvial stages, however, consid- erable connectedness of drainages can be as- sumed. Figure 1 shows maximum stands and fluvial connections of many of the larger Pleistocene lakes of the region, in addition to possible Pliocene coverage of Lake Idaho. Depicted connections and high stands were not all contemporaneous. Little is known of early and middle Pleistocene pluvials. Older pre-Bonneville and pre-Lahontan lacustrine sediments are correlated with Cedar Ridge (Kansan) gla- ciation in the Rocky Mountains and are overlain by the Pearlette (restricted, type-0) Fig. 1. Noncontemporaneous maximum extent of Late Pleistocene lakes and known fluvial connections in intermountain western North America and possible maximum extent of Pliocene Lake Idaho. Compiled and modified from Miller (1946a), Hubbs and Miller (1948), Trimball and Carr (1961). Feth (1961), Bright (193), Snyder et al. (1964), Morrison (1965), and Hubbs et al. (1974). 20 GREAT BASIN NATURALIST MEMOIRS No. 2 ash, which is 600,000 years old (Morrison 1965, Richmond 1970). Well-developed soils separate the older pre-Bonneville and pre- Lahontan sediments from younger pre-Bon- neville and pre-Lahontan lacustrine sedi- ments, believed to be contemporaneous with the Sacagawea Ridge (Illinoian) glacia- tion. These units are also overlain by sub- aerial units and soils believed correlated with the last great interglacial (Sangamon), for which dates are around 130,000 years BP (Richmond 1970). The Alpine formation of the Bonneville Basin and Aetza formation of the Lahontan Basin are thick units of lacustrine deposits frequently interrupted by soils and subaerial zones. They appear to span a time begin- ning perhaps as early as 70,000 years BP, correlative with the Bull Lake glaciation. The latter portion of this period was appar- ently a time of intermediate lacustrine oc- cupation of other pluvial basins as well, e.g., Searles (Smith 1968) and Yellowstone (Birkeland et al. 1971). The period from about 35,000 to 25,000 was apparently dominated by widespread subaerial deposi- tion and soil formation (Morrison 1965, Bir- keland et al. 1971). The most recent pluvial episode, corre- lated with the Pinedale glaciation, has left the most abundant and clear record (e.g., probably most of the Great Basin lakes shown in Figure 1). Radiocarbon dates for high and low stages of five lakes appear to be only partly synchronous. Data for Bonne- ville (Morrison 1965, Bright 1966), Lahon- tan (Broecker and Kaufman 1965, Morrison 1965), Searles (Smith 1968), Mojave (Ore and Warren 1971), and Yellowstone (Rich- mond 1970) agree in showing development of generally, but not consistently, high lake levels during the period between 25,000 and 13,000 years BP. In each lake, a low stand is marked at about 11,000 years, fol- lowing a high stand in the previous one or two thousand years. High levels are record- ed again in the interval between 11,000 and 10,000 years, followed by low stand at about 9,000 years BP and unstable-low stands throughout the warm period of 8,000- 4,000 years BP and to the present. Lake Malheur, in the Harney Basin, shows a broadly similar pattern (Hansen 1947). Lo- cal details will be mentioned below, in con- nection with special distributional problems. Although there is a tendency to look to the 13,000 and 11,000 BP lacustrine highs and their inferred associated climates for ex- planation of distributional patterns, it is im- portant to remember that they represent only an unstable, late episode among a number of intermittent pluvial periods. The role of geologic and climatic factors in the manipulation of barriers and habitats will be examined by analyzing fish distributions to discover the major patterns and possible determinants of the patterns. Methods Of the 70 or more major drainage basins, many are fishless and will not be considered in this study. The basins with fishes are di- vided into barrier-bounded, faunally homo- geneous units. Criteria for barriers are pres- ence of mountains or deserts uncrossed by continuous aquatic habitat or at least one- way restriction of fish dispersal, such as bar- rier falls. Criteria for homogeneity include the restriction that a drainage unit should not have two or more barrier-separated areas of endemism within it. For practical purposes, contiguous but separated small areas were often joined as one if they con- tained the same fauna. These criteria al- lowed the relatively objective designation of 48 drainage units, which are outlined and labeled in Fig. 2. To aid in ease of refer- ence, drainage units are given initials in- dicating general region and faunal affinity: (B) Bonneville, (S) Snake, (O) Oregon lakes (some of which are in Nevada and Califor- nia), (K) Klamath, (G) Goose (L) Lahontan, (R) Ruby group (east-central Nevada), (D) Death Valley (with Owens River and Mo- jave Basin), and (C) Colorado. Subsequent initials in each label refer to the specific area or some valley or ancient lake in it. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 21 In the analyses, the drainage units are clustered and ordinated according to the similarity of their fish faunas. The presence or absence of fish species and higher taxa are the characteristics by which drainage units are evaluated. General dispersal pat- terns are also sought by clustering species according to the similarity of their geogra- phic ranges. The methods used are single linkage cluster analysis (the similarity coefficients used are Jaccard's coefficient and correlation coefficients) and principal components (of the among-species cov- ariance matrix), from the general field of numerical taxonomy (Sneath and Sokal 1973). Variation in species density and its causes are examined by regression of species num- bers on area and by graphs examining spe- cies density in drainage units and breadth of distribution of species among units. Histori- cal perspectives are sought through an ex- Fig. 2. Outline and identification of 48 drainage units as defined in this analysis. BONNEVILLE group, Utah, Idaho, and Nevada: BB— Bear R.-We- ber R. dr., BT-Thousand Springs-northwest Great Salt Lake dr., BP-Provo R.-Utah L.-Jordan R. dr., BD-Deep Cr. dr., BSn-Snake V. and related dr., BSv-Sevier R. dr., BSh-Shoal Cr. (Pine Canyon Cr.) dr.; SNAKE R. group, Idaho, Wyoming, Utah, Nevada, Oregon: SU-Upper Snake above Shoshone Falls, SL-Lost R.-Camas Cr. dr., SW-Wood R. dr., SM-Middle Snake R. (below Shoshone Falls and Wood R. falls); OREGON LAKES group, Oregon, Nevada, California: OH-Harney Basin, OAk-Alkali L. dr., OAb-Abert L. dr., OSm-Summer L. dr., OFR-Fort Rock-Christmas L. Basin, OSi-Silver L. dr., OW-Wamer L. dr., OC-Catlow V. dr., OAl-Alvord dr., OL-Long V. (Massacre L.) dr., OSp-Surprise V (Alkali lakes) dr.; G-GOOSE LAKE, California, Oregon; K-KLAMATH system (upper), Oregon, California; LAHONTAN group, Nevada, California, Oregon: L-Lahontan (Humboldt R., Pyramid L„ L. Tahoe, Walker L., Crescent V., Grass V.), LDm-Diamond V., LE-Eagle L. dr., LMd-Madeline Plains, LDx-Dixie V., LT-Toiyabe (Big Smokey V.), LF-Fish L. dr., LN-Newark V., LC-Clover-Independence V. (Snowater L.). RUBY group, Nevada: RF-Ruby-Franklin dr., RB- Butte V., RW-L Waring (Goshute V.), RSt-Steptoe V.; COLORADO group, Utah, Nevada, Arizona: CRR-Railroad V., CW-White R. (Muddy R.) dr., CM-Meadow Valley Wash dr., CV-Virgin R. dr., C-upper Colorado dr. (above Virgin R.), CLV-Las Vegas dr.; DEATH VALLEY group, Califor- nia, Nevada: DO— Owens dr., DMo— Mojave dr., DMn— Amargosa-Manly dr., DT-Amargosa-Tecopa dr., DP-Pahrump V. 22 GREAT BASIN NATURALIST MEMOIRS No. 2 animation of the fish fossil record for data on past distributions, species density, and amounts of taxonomic change. Fish Fauna The fishes inhabiting the Intermountain Region are generally classified (Bailey et al. 1970) in 83 species, 26 genera, and 7 fami- lies; 46 of the species (55 percent) and 6 of the genera (13 percent) are endemic (Table 1). In the section that follows, the species are listed with a statement of the range within and beyond the study area (abbrevia- tions refer to drainage units in Figure 2), fossil record, and other special information. Fossil species from the region that have liv- ing representatives in the region are includ- ed in the list and marked with a + . Petromyzontidae Lampetra minima Bond and Kan 1973. Miller Lake Lamprey. K (Miller L., Ore.), Apparently extinct. Lampetra tridentata (Gairdner in Richard- son) 1836. Pacific lamprey. SM, K, G; coast- al drainages, Alaska to S Calif., and E Asia. Lampetra lethophaga Hubbs 1971. Pit- Klamath brook lamprey. K (upper Klamath dr., Ore.); Pit R. dr., Calif. Table 1. Numbers of families, genera, and species of native fishes in the Intermountain Region, as considered in this report. Families Petromyzontidae lampreys Genera 1 Species 3 (1 endemic) AC1PENSEBIDAE 1 1 sturgeons Salmonidae trouts, whitefish 4 10 (3 endemic) Cyprimdae minnows 13 (4 endemic) 28 (14 endemic) Catostomidae suckers 3 21 (11 endemic) Cyprinodontidae killifishes 3 (2 endemic) 8 (8 endemic) Cottidae sculpins 1 12 (6 endemic) Acipenseridae Acipenser transmontanus Richardson 1836. White Sturgeon. SM; coastal streams, Alaska to Calif. Salmonidae Salvelinus malma (Walbaum) 1792. Dolly Varden (the interior "bull trout" popu- lations may be a distinct species [Cavender 1969]). SL (Hubbs and Miller 1948b), SM (Miller and Morton 1952), K (Bond 1973), possibly B (Rostlund 1951), but not counted in this study; freshwater and anadromous, NW N America and E Asia. Rare in Inter- mountain Region. Pliocene relative, west central Nevada (Cavender 1969). Salmo clarki Richardson 1836. Cutthroat trout. BB, BP, BSv, BD, BSn, SU, SL, SW, SM, K, G, OA1, L, LE, CV, C; Western North America from headwaters of South Saskatchewan, Missouri, South Platte, Ar- kansas, Pecos, and Rio Grande drainages to Eel R. in N Calif, and to SE Alaska. Much local differentiation. Threatened or extinct over most of former range in Great Basin; genetically modified by introgression from cultured and introduced upper Snake River forms and Salmo gairdneri (Miller 1961, 1977). Pleistocene, L. Bonneville (Smith et al. 1968). Pliocene S. esmeralda LaRivers (1966), Esmeralda Co., Nev., may be a rela- tive. Salmo gairdneri Richardson 1836. Rain- bow trout. SM, K; Pacific drainage, N America, from N Mexico to W Alaska, and probably headwaters of Peace and Ath- abaska drainage. Widely introduced. Salmo sp. Redband trout. SM, G, OH, OAb, OSi, OW, OC; McCloud and Pit dr., Calif. (Bond 1973, LeGendre et al. 1972, LeGendre 1976). + Salmo sp. Pliocene, Lake Idaho (SM), Smith 1975, may be a relative of one or more of the above trouts. Salmo apache Miller 1972. Arizona trout. C (upper L. Colorado dr.); Salt R. dr., Ari- zona. Status: threatened (Miller 1977). Oncorhyncus tshawytscha (Walbaum) 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 23 1792. Chinook Salmon. SM; Pacific dr. from S Calif, to Hokkaido, W Arctic dr. Ana- dromous. Pliocene congener, O. salax Smith 1975, L. Idaho (SM). Prosopium williamsoni (Girard) 1857a. Mountain whitefish. BB, BP, BSv, SU, SL, SW, SM, OH, L, C; northward in head- waters of upper Missouri, Milk, Saskatche- wan, Peace, and Stikine drainages (Scott and Crossman 1973). Prosopium abyssicola (Snyder) 1919, Bear Lake whitefish. BB (Bear L., Utah and Idaho). Prosopium spilonotus (Snyder) 1919. Bon- neville whitefish. BB (Bear L.) Pleistocene, L. Bonneville, Smith et al. 1968. Prosopium gemmiferum (Snyder) 1919. Bonneville cisco. BB (Bear L.) Pleistocene, L. Bonneville, Smith et al. 1968. + Prosopium prolixus Smith 1975, Plioc- ene, L. Idaho (SM). Cyprinidae Acrocheilus alutaceus Agassiz and Picker- ing 1855. Chiselmouth. SM, OH; northward in the Columbia and Fraser drainages. Pliocene relative, A. latos Cope, L. Idaho (SM), (Smith, 1975). Eremichthys acros Hubbs and Miller 1948a. Desert dace. L (Soldier Meadows). Status: rare (Miller 1977). Gila atraria (Girard) 1857b. Utah chub. BB, BT, BP, BSv, BSn, BSh, SU. Extensive local geographic variation in body size, number of gill rakers, color, etc. Pleisto- cene, L. Bonneville, Smith et al. 1967. Gila coerulea (Girard) 1857b. Blue chub. K; Klamath B. Pliocene G. milleri Smith 1975, is a related form from L. Idaho (SM). Gila robusta Baird and Girard 1854a. Boundtail chub. CV, CW, C. Some geo- graphic variation. Gila elegans Baird and Girard 1853. Bo- nytail. C. Status: endangered (Miller 1977). A related form is known from the Pliocene of Arizona (Uyeno and Miller 1965). Gila cypha Miller 1946b. Humpback chub. C. Status: endangered (Miller 1977). Gila (Siphateles) bicolor (Girard) 1857b. Tui chub. K, G, OH, OAb, OSm, OSi, OW, OC, OFB, OAk, L, LCI, LDm, LN, LE, LF, LDx, LT, DO, DMo, CRB. Considerable ge- ographic variation in gill rakers, meristics, osteology, size, etc.; many well-differen- tiated subspecies (Snyder 1917, Miller 1973, Hubbs and Miller 1972, Hubbs et al. 1974). Gila bicolor pectinifer is an ecologically and morphologically divergent form that is ap- parently partially reproductively isolated from sympatric G. b. bicolor under favor- able ecological circumstances in the Lahon- tan Basin (Hubbs et al. 1974, Hopkirk and Behnke 1966). Pliocene Gila turneri (Lucas 1900), ( = G. esmeralda LaRivers 1966), Es- meralda Co., Nev., may be a relative. Gila (Siphateles) alvordensis Hubbs and Miller 1972. Alvord chub. OA1. Gila (Snyderichthys) copei (Jordan and Gilbert) 1880. Leatherside chub. BB, BP, BSv, SU, SW. Hesperoleucus symmetricus (Baird and Gi- rard) 1855. California roach. G; Sacramento dr., coastal streams, Calif. Iotichthys phlegethontis (Cope) 1874. Least chub. BB, BP, BSv, BSn. Status: vul- nerable (Miller 1977). Richardsonius egregius (Girard) 1858. La- hontan redside L,LE. Richardsonius balteatus (Richardson) 1836. Redside shiner. BB, BT, BP, BSv, BSh, SU, SW, SM, OH; northward through the Columbia dr. to the Nass R., B. C, and the Peace R., B. C. and Alberta. The form of the Bonneville, Upper Snake, Palouse, Har- ney Basin (except Silvies R.) and some iso- lated headwaters of the M Snake, is differ- entiated, with fewer rays in the anal fin (R. b. hydrophlox). Richardsonius durranti of Pliocene L. Idaho (SM) is a relative (Smith 1975). Ptychocheilus oregonensis (Richardson) 1836. Northern squawfish. SM, OH; north- ward through Umpqua, Siuslaw, and Colum- bia drainages to the Nass R., B. C, and the Peace R., B. C. and Alberta. Ptychocheilus arciferus of Pliocene L. Idaho (SM) is a rel- ative (Smith 1975). Ptychocheilus lucius Girard 1857b. Colo- 24 GREAT BASIN NATURALIST MEMOIRS No. 2 rado squawfish. C. Status: endangered (Mill- er 1977). P. prelucius, Pliocene of Arizona (C) is a relative (Uyeno and Miller 1965). Mylocheilus ca minus Richardson 1836. Peamouth. SM; northward through the Co- lumbia drainage to the Nass R. and the Peace R., B. C. M. robustus, Pliocene L. Idaho, is a relative (Smith 1975). Rhinichthys cataractae (Valenciennes) 1842. Longnose dace. BB, BP, SU, SW, SM, OH; widespread throughout northern U.S. and southern Canada, north to the Mac- kenzie and south to N Mexico along the Rockies and from Ungava drainage south to Tennessee and North Carolina in the East. Rhinichthys falcatus (Eigenmann and Ei- genmann) 1893. Leopard dace. SM; Colum- bia, Fraser drainages. Rhinichthys osculus (Girard) 1857b. Speckled dace. BB, BT, BP, BSv, BD, BSn, BSh, SU, SW, SM, K, G, OH, OAb, OSi, OW, OFR, OSp, OL, L, LCI, LE, LMd, LDm, LT, DO, DT, CV, CW, CM, CLV, C; elsewhere in Pacific drainage N. A. from the Columbia dr. of extreme southern B. C. (Scott and Crossman 1973) to northwestern Mexico (Sonora), New Mexico, Arizona and Calif. Geographically variable in meristics, color, size, and proportions (Hubbs et al. 1974). Rhinichthys sp. BSn (Bonneville desert). Relictus solitarius Hubbs and Miller 1972. Relict dace. RF, RSt, RW, RB. Moapa coriacea Hubbs and Miller 1948a. Moapa dace. CW Status; endangered (Mill- er 1977). Lepidomeda albivallis Miller and Hubbs 1960. White River spinedace. CW (Preston and Lund Spr., Nevada). Status: vulnerable (Miller 1977). Lepidomeda altivelis Miller and Hubbs 1960. Pahranagat spinedace. CW (Ash Spr. and upper Pahranagat L., Nev.). Status: ex- tinct. Lepidomeda mollispinis Miller and Hubbs 1960. Virgin spinedace. CV, CM (distinct subspecies in Meadow Valley Wash, ex- tinct). Status: vulnerable (Miller 1977). Lepidomeda vittata Cope 1874. Little Colorado spinedace. C. (Little Colorado dr.). Status: vulnerable (Miller 1977). Plagopterus argentissirnus Cope 1874. Woundfin. CV; (and formerly from the Gila R. dr.). Status: endangered (Miller 1977). Catostomidae Catostomus ardens Jordan and Gilbert 1880. Utah sucker. BB, BP, BSv, BD, BSn, SU. Pleistocene Lake Bonneville (Smith et al. 1968). Catostomus macrocheilus Girard 1857b. Largescale sucker. SM, OH; northward through Columbia drainage to Nass R., B.C., and Peace R., B. C. and Alberta. (?) Pliocene and Pleistocene relatives, Lake Idaho (Smith 1975). Catostomus occidentalis Ay res 1854. Sac- ramento sucker. G; Sacramento dr., and coastal drainages, Calif. Catostomus sp. OSp (trib. Surprise Valley, Nev.). Catostomus warnerensis Snyder 1908. Warner sucker. OW. Status: endangered (Miller 1977). Catostomus tahoensis Gill and Jordan in Jordan 1878. Tahoe sucker. L, LE. Catostomus fumeiventris Miller 1973. Owens sucker. DO. Catostomus insignis Baird and Girard 1854b. Sonora sucker. CV (formerly); lower Colorado drainage. Catostomus latipinnis Baird and Girard 1854b. Flannelmouth sucker. CV, C; lower Colorado. Catostomus snyderi Gilbert 1898. Kla- math largescale sucker. K; Klamath dr. Catostomus catostomus (Forster) 1773. Longnose sucker. SU; northern U.S., Can- ada, Alaska, NE Asia. Catostomus (Pantosteus) columhianus (Ei- genmann and Eigenmann) 1893. Bridgelip sucker. SW, SM, OH; Columbia and Fraser drainages. C. arcnatus of Pliocene Lake Idaho is a relative (Smith 1975). Catostomus (Pantosteus) discobolus Cope 1872. Bluehead sucker. BB, SU, C. Catostomus (Pantosteus) clarki Baird and Girard 1854b. Desert sucker. CV, CW, CM; 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 25 lower Colorado drainage. Catostomus (Pantosteus) platyrhynchus (Cope) 1874. Mountain sucker. BB, BP, BSv, BD, BSh, SU, SM, L, C; upper Missouri, Saskatchewan, Columbia, and Fraser drain- ages. Catostomus (Deltistes) luxatus Cope 1879. Lost R. sucker. K; Klamath dr., Ore., Calif. Catostomus owyhee of Pliocene L. Idaho is a relative (Smith 1975). Clxasmistes brevirostris Cope 1879. Short- nose sucker. K; Klamath dr. Ore., Calif. Status: vulnerable (Miller 1977). Chasmistes cujus Cope 1883. Cui-ui. L. Status: endangered (Miller 1977). Chasmistes liorus Jordan 1878. June suck- er. BP. Status: threatened or extinct (Miller 1977). Chasmistes sp. SU. Status: extinct. C. spa- tulifer, Pliocene L. Idaho, is a relative (Smith 1975). + Chasmistes spp. Undescribed Plio- Pleistocene fossil Chasmistes are known from four other localities: Owens Valley dr. (DO), Madeline Plains (OMd), Fossil Lake (OFR), and the Thatcher Basin, Idaho (BB) (Miller 1965; Bright 1967). Xyrauchen texanus (Abbott) 1861. Hump- back sucker. C. Status: vulnerable (Miller 1977). Cyprinodontidae Empetrichthys merriami Gilbert 1893. Ash Meadows killifish. DT. Status: extinct. Empetrichthys latos Miller 1948. Pahrump killifish. DP. Status: endangered (in refuges; extinct in native habitat) (Miller 1977). £. erdisi (Jordan) 1924a from the Pliocene of Ridge Basin, L.A. Co., Calif., is a relative (Uyeno and Miller 1962). Crenichthys baileyi (Gilbert) 1893. White River springfish. CW. Status: special con- cern (Miller 1977). Crenichthys nevadae Hubbs 1932. Rail- road Valley springfish. CRR. Status: special concern (Miller 1977). + Fundulus spp.— Five species of Late Cenozoic Fundulus are known from S Calif, and Nevada (Death Valley, Mojave dr., La- hontan dr., and Ridge Basin); they are rela- tives of Empetrichthys and Crenichthys (Uyeno and Miller 1962). Cyprinodon salinus Miller 1943. Salt Creek pupfish. DMn. Cyprinodon nevadensis Eigenmann and Eigenmann 1889. Amargosa pupfish. DMn, Dt. Geographically variable (Miller 1948, LaBounty and Deacon 1972). Status: several subspecies rare or endangered, one extinct (Miller 1977). Cyprinodon diabolis Wales 1930. Devils Hole pupfish. DT. Status: endangered (Mill- er 1977). Cyprinodon radiosus Miller 1948. Owens pupfish. DO. Status: endangered (Miller 1977). + Cyprinodon breviradius Miller 1945, from the Tertiary of Death Valley is related to the above four species. Cottidae Cottus bairdi Girard 1850. Mottled scul- pin. BB, BT, BP, BSv, BD, BSn, SU, SM, OH, C; Columbia drainage north to British Columbia, east through Missouri, L. Winni- peg, Hudson Bay and Ungava drainages, and across northern U.S. south to Oregon, Nevada, Utah, New Mexico, Montana, Iowa, Missouri, Alabama, and Georgia. Late Pleistocene fossil, L. Bonneville (Smith et al. 1968). Geographically variable (Bisson and Bond 1971, Bond 1963). Cottus confusus Bailey and Bond 1963. Shorthead sculpin. SL, SM; Columbia, Pu- get Sound, and Flathead R. drainages. Cottus extensus Bailey and Bond 1963. Bear L. sculpin. BB (Bear L.). Late Pleisto- cene fossil, L. Bonneville (Smith et al. 1968). Cottus echinatus Bailey and Bond 1963. Utah L. sculpin. BP (Utah L.). Status: ex- tinct. Cottus beldingi Eigenmann and Eigen- mann 1891. Piute sculpin. BB, SU, SL, SM, C (Grand R. dr., Colorado, only; not includ- ed in calculations), L; Columbia dr. Plio- cene fossil, Lahontan basin (Jordan 1924, Hubbs and Miller 1948b). 26 GREAT BASIN NATURALIST MEMOIRS No. 2 Cottus leiopomus Gilbert and Evermann 1894. Wood River sculpin. SW. Cottus greenei (Gilbert and Culver) 1898. Shoshone sculpin. SM (Thousand Springs area, mouth of Salmon Falls Cr., Idaho). Cottus pitensis Bailey and Bond 1963. Pit sculpin. G; Pit R. dr., California and Oregon. May be extinct in Oregon (Bond 1973). C. calcatus Kimmel 1975 of the Miocene-Pliocene Deer Butte fm., SE Ore- gon, (SM), is a relative. Cottus princeps Gilbert 1898. Klamath Lake sculpin. K (Klamath L.). Cottus klamathensis Gilbert 1898. Mar- bled sculpin. K (Klamath basin); upper Pit R., Calif. Cottus tenuis (Evermann and Meek) 1898. Slender sculpin. K (Klamath basin). Cottus rhotheus (Smith) 1883. Torrent sculpin. SM; Columbia dr. and nearby coastal streams, Puget Sound dr., Fraser dr. southern B.C. Significance of the Fossil Record Consideration of the fossil evidence, as noted among the species accounts, leads to several general conclusions that serve as a background for analysis of history of the fauna. Those conclusions are as follows. (1) Many genera of the intermountain fish fauna were in the region, and in some of the areas presently occupied, by Pliocene time (e.g., Prosopium, Salve- linus, Salmo, Oncorhynchus, Acrocheilus, Gila, Mylocheilus, Ptychocheilus, Richard- sonius, Catostomus [3 subgenera], Chas- mistes, Empetrichthys, and Cottus). (2) Pliocene forms are specifically differ- ent from Recent forms. (3) Pleistocene forms are not usually spe- cifically distinguishable from Recent counterparts. (4) The general latitudinal gradients that exist today are also reflected in the fossil occurrences; i.e., Cottus was restricted to the northern regions, cyprinodontids were restricted to the southern regions, except that in pluvial periods some northern forms (e.g., Chasmistes) were much further south, and in the Pliocene some faunas included forms now dis- placed to subarctic regions (e.g., Myo- xocephalus) as well as forms now dis- placed to the south (e.g., Orthodon, Mylopharodon, Archoplites; Lake Idaho, Smith 1975). (5) During some times in the past, distri- butions were broader and faunas larger than at present (corollary of No. 4; e.g., Chasmistes; Lake Idaho fauna). Analysis of Fish Distributions What are the major patterns of distribu- tion? From the basic data of records of na- tive occurrence of intermountain fishes among the 48 drainage units, a matrix was computed showing the correlation of each species distribution with every other within the region. Correspondence of species pat- terns can be examined by similarity indices and by correlation coefficients. In this case, the product-moment correlation coefficient was found to be less sensitive to simple widespread abundance and was chosen as the basis for discussion. The correlations are summarized by a single-linkage phenogram, in which species are clustered together ac- cording to the correlation among their dis- tributions (Fig. 3). The general nature of the phenogram may be described in five general sections, (1) a large, rather closely clustered Kla- math-Snake R. -Bonneville group; (2) a group inhabiting the Colorado Basin; (3) a small group characterizing the Lahontan Basin; (4) Owens Basin and Death Valley forms; and (5) isolated species with patterns unlike any others. The apparent similarity of patterns of Klamath and middle Snake species is exag- gerated somewhat by the fact that many of their species appear only in one or both of those units within the study area, but have diverse distributions outside. Nevertheless, the similarity of patterns among the north- ern drainages, including the Bonneville and 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 27 the Colorado, is much higher than among the southern drainages. Factors contributing to this phenomenon will be discussed below. A second observation is that species of the pluvial Bonneville Basin are scattered in a number of small diverse clusters, sug- gesting an unusually complex history for this assemblage. Lahontan species show the same trend in a less extreme way; here, however, the scattered forms are those widespread species that are found in numer- ous other drainages (e.g., Salmo clarki, Pros- opium williamsoni, Rhinichthys osculus, and Gila bicolor). Relictus solitarius, which inhabits isolated basins of east central Nevada, RF, RSt, RW, and RB, has a unique distribution pattern. Of the 14 species connecting at a level more remote than .5, all but 2, Rhinichthys osculus and Gila bicolor, are restricted to very isolated, depauperate basins. At least one cluster of species comprises widespread, ecologically similar forms, whose dispersal modes and histories might be inferred to be somewhat similar. Cottus bairdi, Prosopium williamsoni, Richardsonius balteatus, Rhinichthys cataractae, and Ca- tostomus platyrhynchus are the nucleus of a Bonneville-Snake medium- and small-stream fauna; some of these, or their vicariants, are also found in the Lahontan and Colorado systems and other northern mountain drain- ages. They are often ecologically associated with Salmo clarki (upstream) and Rhii- nichthys osculus (downstream). It can also be inferred that any habitat that supports the majority of this group might well be ex- pected to support the others. (The success of introduced Richardsonius balteatus in the Green River drainage is an example.) These forms, especially the trout and sculpin, oc- cupy headwaters and as such are subject to a higher-than-average incidence of dispersal by stream capture. (A process by which "dispersal" to a new drainage system occurs without the individual fishes necessarily leaving a limited home range.) If dispersal by stream capture were the primary factor determining the breadth of distribution of these fishes, we should see a correlation be- tween degree of headwater habitat prefer- ence and number of drainages occupied SIMILARITY .9 .8 .7 .6 .5 .4 .3 .2 Colorado R. 6 sp. Catostomus latipinnis Gila robusta Catostomus insignis — Plagop. argentissimus- Catostomus clarki Lepidom. mollispinis • White R. 4 sp. Goose L. 3 sp. Bear L. 4 sp. Chasmistes sp. Ca tos tomus ca tos tomus Cottus leiopomus Klamath 9 sp. ! Lampetra tridentata — Salmo gairdneri Salvelinus malma Cottus confusus Middle Snake R. 6 sp.| Catost. columbianus — — i Catost. macrocheilus — ■ Acrocheilus aJutaceus — M Ptycho. oregonensis — ' Cottus bairdi Prosopium williamsoni Richardson, balteatus Rhinichth. cataractae Catost. plati/rhynch Gila copei Gila atraria Catostomus ardens Idtich. phlegethont Cottus beldingi Ca tos t . di scobol us ■ Salmo clarki Salmo sp. (redband) Rhinichthys sp. Cottus echinatus ■ Chasmistes liorus Chasmistes cujus Eremichthys acros Richardson, egregius Catost. tahoensis Catost. warnerensis Gila bicolor Rhinichthys osculus Gila alvordensis Catost. fumeiventris Cyprinodon radiosus Crenichth. nevadae Empetrichth. merriami Cyprinodon di Cyprinodon nevadens Cyprinodon sa Catostomus sp. Empetrichthys latos Relictus solitarius mi I :us— Jl :ae 'V ius ' 'B^ "-T3 us 2) r k merriami — | abolis — I I vadensis r linus I Fig. 3. Cluster analysis of 83 species of inter- mountain fishes based on correlation of their distribu- tions among 48 drainage units. Species with similar patterns cluster together; species with unique or dis- tinct patterns cluster at remote levels. 28 GREAT BASIN NATURALIST MEMOIRS No. 2 (compare species list and Figure 4). The correlation does not exist, suggesting that breadth of habitat occupiable by a species, and extinction resistance, are about as im- portant as frequency of colonization events. The most widespread species are Rhi- nichthys osculus, 32/48 drainages, and Gila bicolor, 21/48 drainages. These are not par- ticularly vagile or otherwise prone to colo- nization, but are interpreted to be extinc- tion-resistant generalists with ability to persist in small streams and desert spring habitats as well as other aquatic environ- ments. It is inferred that the distribution of these and other intermountain fishes has been shaped by a few successful coloniza- tions and many extinctions. The model of MacArthur and Wilson (1963, 1967) is applicable to the study of a system such as that described in the preced- ing paragraph. Figure 4 shows the relation- ship between species and the number of drainages occupied. Fifty-one of the 83 spe- cies occupy only one drainage unit in the study area. Eighteen of these are endemic to single areas in the study; 33 are more widespread outside the area, generally in the Columbia, Pit, Klamath, or Colorado drainages. The extremely concave curve suggests a nonequilibrium situation with ex- tinction heavily predominating over coloni- zation. It was noted above that longitudinal dis- tribution of species in the north is distinctly broader than that of species in the south. This could result from any of several pro- cesses: (1) extinction may have been more severe in the south, leading to reduced ranges, (2) there may simply be more longi- tudinally continuous aquatic habitat in the north (omitting the Snake and Colorado drainages from the comparison) because of drainage orientations and the general north- ward increase in precipitation /evaporation ratios, or (3) barriers may be more extreme in the south. The fossil record, though in- complete, suggests that extinction of species has been at least as common in the north jF -/ 9 t~~ju: 2 3 4 5 6 7 8 9 10 II 12 13 14 15 16 NUMBER OF DRAINAGES OCCUPIED BY SPECIES -v r¥ ^V Fig 4 Breadth of distribution of 83 species among 48 drainage units. Species are distributed according to the number of units they occupy. Fifty- one species (61 percent) are found in only one unit; 18 of these are endemic to that unit, 33 have additional range outside the study area. The steep concavity of the curve indicates much extinction and limited colonization. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 29 and that the phenomenon exists indepen- dently of extinction, though possibly in- tensified by it. Factor (2) is observable and intuitively obvious as a causal agent in dis- tribution; factor (3) involves a nonobvious principle that also may be important to the understanding of distribution patterns. Be- cause of the well-known interaction among latitude, altitude, and temperature (e.g., Jan- zen 1967), the habitat of cool-stream fishes tends to be at lower altitudes in the north and higher altitudes in the south. Many northern salmonids, suckers, minnows, and sculpins show this pattern (e.g., Fig. 5). The result is that mountains of a given altitude are not as "high" in terms of barrier effects in the south, in that passes with drainage connections at 8000 ft may be occupied and accessible at 40 degrees north but not at 44 degrees north. Similarly, to these species, lowlands and valleys may be barriers in the south, but not in the north. Conversely, to lowland, warmwater fishes such as cyprino- donts, suitable habitats "pinch out" against the hillsides at lower elevations in the north in much the same way that suitable habitats for sculpins, etc., "pinch out" at upper lim- its of mountain aquatic habitat in the south. That cyprinodonts and sculpins, for ex- ample, have been so ecologically and evolu- tionarily limited by this phenomenon in the Intermountain Region, as opposed to east- ern North America, is significant (cyprino- donts range north to Canada and sculpins range south to Alabama in the east; there is almost no latitudinal overlap in the Inter- mountain Region). It could be concluded that in the Inter- mountain Region mountain barrier effects on stream fishes decrease southward, except that an offsetting corollary also exists: the latitude-altitude-temperature effect that 9 -w i 8 •• •:; • Cot t us bl Hfdi 7 • ••• • • • • • • • • • • • t •• • • ST 6 o • • • • • • • X • • • » 5 • • • /. • • • • • V 4 3 <- i • • •• • • • • z < C o 5 • O Q. z 3 o > PANG 9 o at • • 1 LU 2 " a. • • 1 . • • 6 1 1 CO L-£ • • 1 38° 39° 0° 41° 4 2° 4 3° 44° 45° 46° 4 7 °N Latitude 5. Ecological gradient in habitat of 99 samples of Cordis bairdi from low elevations in the north to high elevations in the south 48° 30 GREAT BASIN NATURALIST MEMOIRS No. 2 brings temperature optima high into the mountains in the south also brings arid desert conditions and higher climatic varia- bility to the tops of many ranges, thus pro- viding the ultimate barrier because aquatic habitat is eliminated entirely. These are places where an observer can stand beneath dark clouds and see rain falling but not reaching the ground. Analysis of Drainage Units The 48 drainage imits will be compared from the standpoints of numbers of species present, similarity (shared species) of the faunas of pairs of units, and strengths of barriers among units. The basic zoogeo- graphic data have been accurately known since the C. L. Hubbs expeditions of the 1930s and 1940s, though different tax- onomic or drainage interpretations change the numbers slightly and, in a few cases, it is not certainly known whether species are indigenous or introduced (Hubbs and Miller 1948b, Hubbs et al. 1974). It is clear from Figures 6 and 7 that northern drainages have more species, on the average. Also, large basins support more species than small basins (Fig. 8). Peripheral areas have more species on the average than those of the in- Fig. 6. Map of drainages showing major existing fluvial and lacustrine habitats, basin outlines as defined in this study, numbers of species of native fishes in each drainage unit large, central numerals), and numbers of species of fishes limited to only one or the other side of each drainage divide bar- rier (small numerals Deal drainage divides). Breaks in dividing lines indicate probable late-pluvial connections (compare Fig. 1). See Fig. 2 for names of drainage units. Basins with no native fishes are unlabeled. Basins with single species are marked with a subscript to the numeral one indicating the in- - Rfunichthys oscultu. s - Hilutw, solitarius. b - Gila bicolor, E - Empetrichthys latos. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 31 terior, probably as a result of their proxim- ity to colonization source areas (see Mac- Arthur and Wilson 1967), but also because areas associated with the Sierra Nevada and Wasatch mountains have higher precipi- tation/evaporation ratios and more aquatic habitat. The pattern of species density among drainage areas (Figs. 4 and 7) is indicative of the degree of isolation and the frequency of extinction among Great Basin fishes. Brown (1971, 1978) showed that the distri- bution of boreal mammals on Great Basin mountaintop islands was characterized by fewer species, especially on small islands, than expected on the basis of the theory of extinction-colonization equilibrium. He fur- ther suggested (1971:477) that the fishes, like the mammals, colonized extensively during pluvial (and boreal flora) maxima, with subsequent isolation, reduced coloniza- tion, and increased extinction rates. A graph of log number of species of fishes against log drainage area (Fig. 8) confirms Brown's suggestion and provides an even more ex- treme example. The z (slope) value for the fish example is 0.59 (r = .66). This value is considerably higher than the value (z = .43) for Great Basin mammals, indicating an extremely high extinction rate (see also Hubbs et al. 1974:79). The paucity of spe- cies in the majority of basins, especially compared with potential source areas (Figs. 6 and 7), is evidence for an extremely low colonization rate. The strengths of the barriers among ba- sins is shown by the small numerals next to barriers in Figure 6. The numerals give the number of species whose distribution does not cross the barrier. For example, 21 spe- cies of fishes are restricted above or below the falls of the Snake R. in units SU and and SM; of the 14 species in the Colorado drainage (C) and the 13 in the Utah Lake drainage (BP), 17 species fail to cross the Wasatch Mountain barrier. (The number in 15 14 13 12 I I 2 "o .6, except where extinction has been severe but differential.) Barrier strength between the Bonneville and Lahon- tan Basins is high (B = .92), but artificially elevated by severe extinction in the north- west Bonneville Thousand Springs unit (BT). Barriers in the Mojave-Owens-Death Valley systems are strong (B = .75-1.0) because of the high level of taxonomic differentiation in this system (Miller 1948) and because of severe extinction. 17 13 10 uj 7 - o UJ Q- c (Column Totals) 4 3 2 1 0 40 108 67 55 43 49 10 2 3 1 1 1 1 6 2 2 2 5 1 6 18 8 7 7 7 1 25 37 25 17 14 23 4 52 32 29 20 13 3 1 2 3 4 5 6 Number of Native Species per Collection 7 18 54 145 150 (Row Totals) 38 GREAT BASIN NATURALIST MEMOIRS No. 2 netically related forms have a slightly high- er tendency to cluster near each other, partly because they tend to link through their species-group or genus if they do not have a more similar species distribution with which to cluster. Otherwise, the phenograms are identical. Without the phy- logenetic information in the matrix, the re- sults are more ecological and less evolution- ary. Principal components analyses were used because the results are readily interpretable and reveal meaningful structure. Theo- retically the method is inappropriate with categorical data, but the disadvantage is in- significant, compared to the insights provid- ed by the method. Components were calcu- lated from the correlation and covariance matrices. The latter differ from Figure 9 in that the structure was summarized in small- er steps: the first six components accounted for 52 percent of the total information (27 percent in the first two), in ordinations that separately discriminated different drainages on the basis of the uniqueness of their faunas. The components of the correlation matrix were broader in this summarization, revealing more general trends as shown in Figure 9 (the first six components account for 66 percent of the total information). Use of these quantitative methods is justi- fied by the summarizations they afford and by the otherwise nonobvious insights that they provide. Examples of the latter are the discovery of the basic northern fluvial faun- al unit and its relation to other species pat- terns (Fig. 3) and to the drainage units (dis- cussion of Fig. 9). A second example is the factoring out and display of the extinction and dispersal effects by distortion of posi- tions of the units in Figure 9 relative to their positions in geographic space. The barrier analysis (Figure 6) quantified a basic (and intuitively obvious) distinction between postpluvial desiccation barriers and mountain barriers, though severe extinction in some units artificially increased the cal- culated barrier strength. The high barrier indices for the Wasatch Range and the southern edge of the Snake River Plain mark zones of extreme differentiation that are also zones of tectonic interest in that they are the boundaries of the Great Rasin subplate (Smith and Sbar 1974). A second zone of great barrier strength and extreme faunal differentiation is that running north- south in eastern Nevada from the Ruby Mountains to the White River drainage and Meadow Valley Wash. This zone is near the region that may mark the center from which tectonism leading to basin and range structure began and radiated outward (Arm- strong et al. 1969, Scholz et al. 1971). The principal components and barrier analyses also emphasized the distinction between the faunas of the Virgin, Meadow Valley Wash, and White River drainages and that of the rest of the Colorado R., particularly that above Grand Canyon. Origin of the dis- tinction very likely dates from the Pliocene drainage pattern associated with the inter- ruption of the Colorado fluvial system by Hualpai Lake near the mouth of the present Colorado River Canyon in the Grand Wash Cliffs (Hunt 1956, 1969). The problem of the Colorado River con- nections to the Mojave Desert is still un- solved (Hubbs and Miller 1948b) but the fact that the fauna is limited to fishes with brackish water tolerance (except for the three species from the north) suggests that the timing of the connections could date back to the Rouse embayment, some time in the Pliocene (Metzger et al. 1973:34). This suggestion is based on the assumption that a connection after recession of the embay- ment and the establishment of the Colorado River would have allowed colonization by lower Colorado River primary freshwater fishes as well as brackish water forms. Finally, this analysis has proved to be a study of patterns of extinction as well as patterns of dispersal. The results indicate that the two processes are not easily sepa- rated analytically, and that studies of dis- persal based strictly on cladism, ignoring ex- tinction, are in peril of error due to missing faunas and missing sister groups. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM Unfortunately the process of extinction is not restricted to the history of the present example, but is accelerating owing to land misuse and water use conflicts and is avail- able for study as a dynamic process. Intense attention is currently focused on protection of a few individual species, but attention must soon shift to long-term maintenance of stability of aquatic habitats through man- agement of surface waters and ground wa- ters at the level of watershed ecosystems. Greater restraint in the introduction of exotic species would also enhance the survi- val of native species. Acknowledgments Robert R. Miller provided much useful information and helpfully reviewed the manuscript. Ruth L. Elder prepared the maps and Kama Steelquist drafted graphs and photographed the figures. Beverly Smith typed the manuscript. Dwight W. Taylor provided helpful and thought-pro- voking discussion. Literature Cited Abbott, C. C. 1861. Descriptions of four new species of North American Cyprinidae. Proc. Acad. Nat. Sci. Philadelphia 12: 473-474. Agassiz, J. L. R. 1855. Synopsis of the iehthyological fauna of the Pacific slope of North America, chiefly from the collections made by the expe- dition under the command of Capt. C. Wilkes, with recent additions and comparisons with eastern types. Amer. J. Sci. 19: 71-99, 215-231. Armstrong, R. L., E. B. Ekben, E. H. McKee, and D. C. Noble. 1969. Space-time relations of Ce- nozoic silicic volcanism in the Great Basin of the western United States. Amer. J. Sci. 267: 478-490. Abmstrong, R. L., W. P. Leeman, and H. E Malde. 1975. K-Ar dating, Quaternary and Neogene volcanic rocks of the Snake River Plain, Idaho. Amer. J. Sci. 275: 225-251. Atwater, T. 1970. Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Bull. Geol. Soc. Amer. 81: 3513-3536. Axelrod, D. I. 1950. Evolution of desert vegetation in western North America. Publ. Carnegie Inst. Wash. 590: 217-306. Aybes, W. O. 1854. [Description of new fishes from California.] Proc. Calif. Acad. Sci. 1: 3-22. Bailey, R. M., J. E. Fitch, E. S. Herald, E. A. Lachner, C. C. Lindsey, C. R. Robins, and W. B. Scott. 1970. A list of common and scien- tific names of fishes from the United States and Canada. Amer. Fish. Soc. Spec. Publ. 6: 1-150. Bailey, R. M., and C. E. Bond. 1963. Four new species of freshwater sculpins, genus Cottus, from western North America. Occas. Pap. Mus. Zool. Univ. Michigan 634: 1-27. Baird, S. F., and C. Girard. 1854a. Descriptions of some new fishes from the River Zuni. Proc. Acad. Nat. Sci. Philadelphia 1853: 368-369. 1854b. Descriptions of new species of fishes collected by Mr. John H. Clark on the U.S. and Mexican Boundary Survey, under the direction of Lt.-Col. Jas. D. Graham. Proc. Acad. Nat. Sci. Philadelphia 1853: 387-390. 1855. [Cyprinidae of Heermann's collection.] Proc. Acad. Nat. Sci. Philadelphia 7: 135-138. BlRKELAND, P. W., D. R. CRANDELL, AND G. M. Richmond. 1971. Status of correlation of Quaternary stratigraphic units in the western conterminous United States. Quaternary Res. 1: 208-227. Bisson, P. A., and C. E. Bond. 1971. Origin and distribution of the fishes of Harney Basin, Ore- gon. Copeia 2: 268-281. Bond, C. E. 1963. Distribution and ecology of fresh- water sculpins, genus Cottus, in Oregon. Un- published doctoral dissertation, University of Michigan Library, Ann Arbor. 1973. Keys to Oregon freshwater fishes. Ore- gon State Univ. Agric. Exp. Sta., Tech. Bull. 58: 1-42. Bond, C. E., and T. T. Kan. 1973. Lampetra (En- tosphenus) minima, n. sp., a dwarfed parasitic lamprey from Oregon. Copeia 3: 568-574. Bright, R. C. 1963. Pleistocene lakes Thatcher and Bonneville, southeastern Idaho. Unpublished doctoral dissertation, University of Minnesota Library, Minneapolis. 1966. Pollen and seed stratigraphy of Swan Lake, southeastern Idaho: Its relation to region- al vegetational history and to Lake Bonneville history. Tebiwa 9(2): 1-47. 1967. Late-Pleistocene stratigraphy in That- cher Basin, southeastern Idaho. Tebiwa 10(1): 1-7. Broecker, W. S., and A. Kaufman. 1965. Radiocarbon chronology of Lake Lahontan and Lake Bonneville II, Great Basin. Bull. Geol. Soc. Amer. 76: 537-566. Brown, J. H. 1971. Mammals on mountaintops: Nonequilibrium insular biogeography. Amer. Naturalist 105: 467-478. 40 GREAT BASIN NATURALIST MEMOIRS No. 2 1978. The theory of insular biogeography and the distribution of boreal birds and mam- mals. Great Basin Naturalist Mem. 2: 209-227. Cavender, T. M. 1969. An early Pliocene Salvelinus from Nevada with comparative notes on Recent species. Abstracts, American Society of Ich- thyologists and Herpetologists 49th annual meeting, pp. 29-30. Cope, E. D. 1872. Recent reptiles and fishes, pp. 432-443. In: F. V. Hayden, Preliminary report of the U.S. Geological Survey of Wyoming and portions of contiguous territories, part 4. U.S. Government Printing Office, Washington, D.C. 1874. On the Plagopterinae and the ichthyol- ogy of Utah. Proc. Amer. Philos. Soc. 14: 129- 139. 1879. The fishes of Klamath Lake, Oregon. Amer. Naturalist 13: 784-785. 1883. On the fishes of the Recent and Plio- cene lakes of the western part of the Great Ba- sin, and of the Idaho Pliocene lake. Proc. Acad. Nat. Sci. Philadelphia 1883: 134-166. Cottam, W. P. 1961. Our renewable wild lands: A challenge. University of Utah Press, Salt Lake City. Eaton, G. P., R. L. Christiansen, H. M. Iyer, A. M. Pitt, D. R. Mabey, H. R. Blank, Jr., I. Zietz, and M. E. Gettings. 1975. Magma beneath Yellowstone National Park. Science 188: 787- 796. ElGENMANN, C. H., AND R. S. ElGENMANN. 1889. Description of a new species of Cyprino- don. Proc. Calif. Acad. Sci., Ser. II. 17: 270. 1891. Cottus beldingi, sp. nov. Amer. Natu- ralist 25: 1132-1133. Evermann, B. W. 1897. A report upon salmon in- vestigations in the headwaters of the Columbia River, in the state of Idaho, in 1895, together with notes on fishes observed in that state in 1894 and 1895. Bull. U.S. Fish Commiss. 16: 149-202. Evermann, B. W., and S. E. Meek. 1898. A report upon salmon investigations in the Columbia River basin and elsewhere on the Pacific Coast in 1896. Bull. U.S. Fish Commiss. 17: 15-84. Feth, J. H. 1961. A new map of western con- terminous United States showing the maximum known or inferred extent of Pleistocene lakes. U.S. Geol. Surv. Prof. Pap. 424-B: 110-112. Forster, J. R. 1773. An account of some curious fishes, sent from Hudson's Bay; in a letter to Thomas Pennant, Esq. Philos. Trans. 63: 149- 160. Gilbert, C. H. 1893. Report on the fishes of the Death Valley Expedition collected in southern California and Nevada in 1891, with descrip- tions of new species. N. Amer. Fauna 7:229- 234. 1898. The fishes of the Klamath River Basin. Bull. U.S. Fish Commiss. 17: 1-13. Gilbert, C. H., and B. W. Evermann. 1894. A re- port upon investigations in the Columbia River Basin, with descriptions of four new species of fishes. Bull. U.S. Fish Commiss. 14: 169-204. Gilbert, G. K. 1890. Lake Bonneville. U.S. Geol. Surv. Monogr. 1:1-438. Girard, C. 1850. A monograph of the freshwater Cotti. Proc. Amer. Assoc. Advanc. Sci. 1849: 409. 1857a. Contributions to the ichthyology of the western coast of the United States, from specimens in the museum of the Smithsonian Institution. Proc. Acad. Nat. Sci. Philadelphia 8: 131-137. 1857b. Researches upon the cyprinoid fishes inhabiting the fresh waters of the United States west of the Mississippi Valley, from specimens in the museum of the Smithsonian Institution. Proc. Acad. Nat. Sci. Philadelphia 8:165-213. 1858. Fishes. Vol. 10, part IV: 1-400. Report of explorations and surveys to ascertain the most practicable and economical route for a railroad from the Mississippi River to the Pacif- ic Ocean. House of Representatives, 33rd Con- gress, 2nd Session, Ex. Doc. No. 91 (1859). Hansen, H. P. 1947. Postglacial forest succession, climate, and chronology of Pacific Northwest. Trans. Amer. Philos. Soc. 37: 1-130. Holt, R. D. 1960. Comparative morphometry of mountain whitefish. Prosopium williamsoni. Copeia 1960(3): 192-200. Hopkirk, J. D., and R. J. Behnke. 1966. Additions to the known native fish fauna of Nevada. Copeia 1966(1): 134-136. Hubbs, C. L. 1932. Studies of the fishes of the order Cyprinodontes. XII. A new genus related to Empetrichthys. Occas. Pap. Mus. Zool. Univ. Michigan 252: 1-5. 1971. Lamptetra (Entosphenus) lethophaga, new species, the nonparasitic derivative of the Pacific lamprey. Trans. San Diego Soc. Nat. Hist. 16: 125-163. Hubbs, C. L., and R. R. Miller. 1948a. Two new, relict genera of cyprinid fishes from Nevada. Occas. Pap. Mus. Zool. Univ. Michigan 507: 1- 30. 1948b. The zoological evidence: Correlation between fish distribution and hydrographic his- tory in the desert basins of western United States. In: The Great Basin, with emphasis on glacial and postglacial times. Bull. Univ. Utah, Biol. Ser. 10: 17-166. 1972. Diagnoses of new cyprinid fishes of isolated waters in the Great Basin of western 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 41 North America. Trans. San Diego Soc. Nat. Hist. 17: 101-106. HUBBS, C. L., R. R. MlLLEB, AND L. C. Hubbs. 1974. Hydrographic history and relict fishes of the north-central Great Rasin. Mem. Calif. Acad. Sci. 7: 1-259. Hunt, C. R. 1956. Cenozoic geology of the Colo- rado Plateau. U.S. Geol. Surv. Prof. Pap. 279: 1-99. 1969. Geologic history of the Colorado River. U.S. Geol. Surv. Prof. Pap. 669-C: 59-130. Janzen, D. H. 1967. Why mountain passes are high- er in the tropics. Amer. Naturalist 101: 233-249. Jordan, D. S. 1878. Contribution to North American ichthyology. III.R. A synopsis of the family Ca- tostomidae. Rull. U.S. Natl. Mus. 12: 97-237. 1924a. Miocene fishes from southern Califor- nia. Rull. S. Calif. Acad. Sci. 23: 42-50. 1924b. Description of a recently discovered sculpin from Nevada regarded as Cottus bel- dingi. Proc. U.S. Natl. Mus. 65(6): 1-2. Jordan, D. S., and R. W. Evermann. 1896-1900. The fishes of north and middle America. Rull. U.S. Natl. Mus. 47: 1-3313. Jordan," D. S., and C. H. Gilbert. 1880. Notes on a collection of fishes from Utah Lake. Proc. U.S. Natl. Mus. 3: 459-465. Kimmel, P. G. 1975. Fishes of the Miocene-Pliocene Deer Rutte formation, southeastern Oregon. Univ. Michigan Pap. Paleontol. 14: 69-87. LaRounty, J. F., and J. E. Deacon. 1972. Cyprinodon milleri, a new species of pupfish (family Cyprinodontidae) from Death Valley, California. Copeia 4: 769-780. LaRivers, I. 1962. Fishes and fisheries of Nevada. Nevada State Fish and Game Commission, Car- son City. 1966. Paleontological miscellanei. Occas. Pap. Riol. Soc. Nevada 11: 1-8. Legendre, P. 1976. An appropriate space for clus- tering selected groups of western Northern American Salmo, Syst. Zool. 25: 13-195. Legendre, P., C. R. Schreck, and R. J. Rehnke. 1972. Taximetric analysis of selected groups of western North American Sahno with respect to phylogenetic divergences. Syst. Zool. 21: 292-307. Lucas, F. A. 1900. A new fossil cyprinoid, Leu- ciscus turned, from the Miocene of Nevada. Proc. U.S. Natl. Mus. 23: 333-334. MacArthur, R. H., and E. O. Wilson. 1963. An equilibrium theory of insular zoogeography. Ev- olution 17: 373-387. 1967. The theory of island biogeography. Princeton University Press, Princeton, New Jer- sey. Malde, H. E. 1968. The catastrophic late Pleisto- cene Ronneville flood in the Snake River Plain, Idaho. U.S. Geol. Surv. Prof. Pap. 596: 1-52. METZGER, D. G., O. J. LOELTZ, AND R. Irelna. 1973. Geohydrology of the Parker- Rlythe-Cibola area, Arizona and California. U.S. Geol. Surv. Prof. Pap. 486-G: 1-130. Miller, R. R. 1943. Cyprinodon salinus, a new spe- cies of fish from Death Valley, California. Copeia 1943: 68-78. 1945. Four new species of fossil cyprinodont fishes from eastern California. J. Wash. Acad. Sci. 35: 315-321. 1946. Correlation between fish distribution and Pleistocene hydrography in eastern Califor- nia and southwestern Nevada, with a map of the Pleistocene waters. J. Geol. 54: 43-53. 1948. The cyprinodont fishes of the Death Valley system of eastern California and south- western Nevada. Misc. Publ. Mus. Zool. Univ. Michigan 68: 1-155. 1958. Origin and affinities of the freshwater fish fauna of western North America, pp. 187- 222. In: C. L. Hubbs (ed.), Zoogeography. Publ. Amer. Assoc. Advanc. Sci. 51: 1-509. 1961. Man and the changing fish fauna of the American southwest. Pap. Michigan Acad. Sci. 46: 365-404. 1965. Quaternary freshwater fishes of North America, pp. 569-581. In: H. E. Wright, Jr., and D. G. Frey (eds.), The Quaternary of the United States. Princeton University Press, Princeton, New Jersey. 1972. Classification of the native trouts of Arizona with the description of a new species, Salmo apache. Copeia 3: 401-422. 1973. Two new fishes, Gila bicolor snyderi and Catostomus fumeiventris, from the Owens River Rasin, California. Occas. Pap. Mus. Zool. Univ. Michigan 667: 1-19. 1977. Freshwater fishes. Red Data Rook. Vol. 4, Pisces. International Union for the Con- servation of Nature and Natural Resources, Sur- vival Service Commission, Paris. Miller, R. R., and C. L. Hubbs. 1960. The spiny- rayed cyprinid fishes (Plagopterini) of the Colo- rado River system. Misc. Publ. Mus. Zool. Univ. Michigan 115: 1-39. Miller, R. R., and W. M. Morton. 1952. First record of the dolly varden, Salvelinus rnalma from Nevada. Copeia 1952: 207-208. Morrison, R. B. 1965. Quaternary geology of the Great Basin, pp. 265-285. In: H. E. Wright, Jr., and D. G. Frey (eds.), The Quaternary of the United States. Princeton University Press, Princeton, New Jersey. Ore, H. T., and C. N. Warren. 1971. Late Pleisto- cene-Early Holocene geomorphic history of 42 GREAT BASIN NATURALIST MEMOIRS No. 2 Lake Mojave, California. Bull. Geol. Soc. Amer. 82:2553-2562. Richardson, J. 1836. Fauna boreali-americana; or, the zoology of the northern parts of British America, containing descriptions of the objects of natural history collected on the late northern land expeditions, under the command of Sir John Franklin, R. N., vol. 4. Richmond, G. M. 1970. Comparison of the Qua- ternary stratigraphy of the Alps and Rocky Mountains. Quaternary Res. 1: 3-28. Rostlund, E. 1951. Dolly varden, Salvelinus malma (Walbaum)? Copeia 4: 296. Russell, I. C. 1885. Geological history of Lake La- hontan, a Quaternary lake of northwestern Ne- vada. U.S. Geol. Surv. Monogr. 1-288. Scholz, C. H., M. Barazanci, and M. L. Sbar. 1971. Late Cenozoic evolution of the Great Basin, western United States, as an en- sialic interarc basin. Bull. Geol. Soc. Amer. 82: 2979-2990. Smith, G. I. 1968. Late Quaternary geologic and climatic history of Searles Lake, southeastern California, pp. 93-310. In: R. B. Morrison, and H. E. Wright, Jr. (eds.), Means of correlating Quaternary succession. University of Utah Press, Salt Lake City. Smith, G. R. 1966. Distribution and evolution of the North American catostomid fishes of the subgenus Pantosteus, genus Catostomus. Misc. Publ. Mus. Zool. Univ. Michigan 129: 1-132. 1975. Fishes of the Pliocene Glenns Ferry formation, southwest Idaho. Univ. Michigan Pap. Paleontol. 14: 1-68. Smith, G. R., and R. K. Koehn. 1969. Phenetic and cladistic studies of biochemical and morpho- logical characteristics of Catostomus. Syst. Zool. 20: 282-297. Smith, G. R., W. L. Stokes, and K. F. Horn. 1968. Some late Pleistocene fishes of Lake Bonneville. Copeia 4: 807-816. Smith, R. B., and M. L. Sbar. 1974. Contemporary tectonics and seismicity of the western United States with emphasis on the intermountain seis- mic belt. Bull. Geol. Soc. Amer. 85: 1205-1218. Smith, R. 1883. Description of a new species of Vranidea (U. rhothea) from Spokane R., Wash- ington Territory. Proc. U.S. Natl. Mus. 6: 347- 348. Sneath, P. H. A., and R. R. Sokal. 1973. Numerical taxonomy; the principles and prac- tices of numerical classification. W. F. Free- man, San Francisco. Snyder, C. T., G. Hardman, and F. F. Zdenek. 1964. Pleistocene lakes in the Great Basin. U.S. Geol. Surv. Misc. Geol. Inv. Map I- 416. Snyder, J. O. 1908. Relationships of the fish fauna of the lakes of southeastern Oregon, Bull. U.S. Bur. Fisheries 27: 69-102. 1917. The fishes of the Lahontan system of Nevada and northeastern California. Bull. U.S. Bur. Fisheries 35: 31-86. 1919. Three new whitefishes from Bear Lake, Idaho and Utah. Bull. U.S. Bur. Fisheries 36: 1-9. Taylor, D. W. 1960. Distribution of the freshwater clam Pisidium ultrarrwntanwn; a zoogeographic inquiry. Amer. J. Sci. 258A: 325-334. Trimble, D. E., and W. J. Carr. 1961. Late Qua- ternary history of the Snake River in the Ameri- can Falls region, Idaho. Bull. Geol. Soc. Amer. 72: 1739-1748. Uyeno, T., and R. R. Miller. 1965. Middle Plio- cene cyprinid fishes from the Bidahochi forma- tion, Arizona. Copeia 1965: 28-41. Valenciennes, A. 1842. In G. Cuvier, and A. Va- lenciennes (eds.), Histoire naturelle des poissons. 16: 1-472. Walbaum, J. J. 1792. Petri Artedi renovati, Pars, i- v. Ichthyologica: Genera Piscium, part iii. Bib- liotheca et philosophia ichthyologica cura Jo- hannis Julii Walbaumii edidit. Wales, J. H. 1930. Biometrical studies of some races of cyprinodont fishes from the Death Val- ley region, with description of Cyprinodon dia- bolis, n. sp. Copeia 1930: 61-70. ZOOGEOGRAPHY OF REPTILES AND AMPHIBIANS IN THE INTERMOUNTAIN REGION Wilmer W. Tanner1 Abstract.— Few, if any, amphibians and reptiles are endemic to Utah. This is also true for much of the Great Basin, upper Colorado Plateau, southern Idaho, and Wyoming. Many species that would seemingly survive in this inland, mountainous area are not here. Only one widespread salamander and a few frogs and toads have occu- pied suitable habitats in the area. Lizards and snakes, like the amphibians, provide few distributions that extend throughout the area. A migration which presumably followed the Pleistocene Ice Age brought most of the species into the area as climatic conditions warmed. Distribution maps of our modern species and subspecies indicate rather clearly that these vertebrates have in- vaded the Intermountain Region in relatively recent geological time. Only the periphery of Utah and adjoining states to the east and west have been penetrated by many of the species in the regional fauna. Few, if any, of the amphibians and rep- tiles now present in Utah are endemic. This is perhaps also true for all intermountain states except those in the south. There is evidence that for a period of time at the close of the Pleistocene the southern Great Basin deserts were more hu- mid than at present. Studies of fossil pack rat middens (Neotoma lepida) by Wells and Jorgensen (1964), indicate that the low desert "ranges in the vicinity of Frenchman Flat (Nevada Test Site) were significantly less arid than at present. Middens now found in areas where the dominant desert shrubs are Larrea and Coleogyne have the leaves, seed, and twigs of Juniperus os- teosperma imbedded in the crystalline urine." Wells and Jorgensen (1964) suggest that the present zonal position of the pi- nyon-juniper forest in southern Nevada is about 600 meters above its position of about 10,000 years ago. The desert valleys of southern New Mexi- co, Arizona, and California must have been considerably more moist and humid during at least part of each year at about the same geological time as the valley floors and low foothills of southern Nevada were covered by forests of pinyon and juniper. We as- sume that climatic conditions then existed which permitted considerable movement of both reptiles and amphibians. Ballinger and Tinkle (1972) and Larsen and Tanner (1975) have assumed that there were Pleistocene refugia in the southern deserts of the United States and /or Mexico which maintained the ancestral stock from which many of the present species and sub- species of intermountain amphibian and reptilian fauna have arisen. The disjunct populations scattered throughout the "is- land" mountains of New Mexico and Ari- zona are highly suggestive of widespread populations being forced from the low val- leys into cooler, moister mountains as post- Pleistocene drying slowly but continuously changed the valleys into uninhabitable deserts for many species. The xeric condi- tions were, however, an invitation for other species to move in, so that the lower Sono- ran valleys and their associated mountain refugia now support a rich and varied series of amphibians and reptiles. If we accept the hypothesis that there was a period of time between the cold, wet Pleistocene and the dry hot conditions of 'Life Science Museum, Brigham Young University, Provo, Utah 84602. 43 44 GREAT BASIN NATURALIST MEMOIRS No. 2 today when much of the southwestern United States was warm but still more hu- mid and moist than at present, we can envi- sion a time in which a great migration of amphibians and reptiles moved toward the plateaus and the mountainous areas of the west central United States. Nevertheless, many species of the south did not penetrate to environments in the Intermountain Re- gion which we might expect to be compatible with their needs. Many North American species that would seemingly survive today in intermountain environments are not here. In June 1942 I had the privilege of escorting Dr. A. H. z j \ ~~~\r r~~S_ \-Js0at\t^ -~L\i ttZ 1 I <^L \r^\ ft jSBISl a Fig. 1. (a) Known western fossil records of Oph- isaurus attenuates; (b) isolated populations of two plethodontid salamanders: upper, Plethodon neo- mexicanus, lower Aneides hardyi. Wright into various habitats in central Utah. He was indeed disappointed not to find salamanders in the debris and leaf mold in and adjacent to mountain springs. Some species of Plethodontidae have reached northern Idaho and the mountains of central New Mexico. Those in Idaho are related to species along the coast from Washington south into California. We can only surmise that the ancestral stock of these species in- vaded our area from north central North America after the Ice Age while the north- ern tier of states was moist enough to per- mit their movement. It is puzzling why some did not persist in the Snake River Val- ley and the Uinta and Wasatch mountains of Utah. At the close of the Pleistocene, some spe- cies expanded their ranges westward across the Great Plains. One example is cited by Etheridge (1961) in which fossil remains of Ophisaurus attenuatus were found in glacial deposits some distance west of their present range (Fig. la). Unfortunately, few fossils are available to substantiate the movement of other species. The disjunct ranges of many species are highly suggestive however, of the westward movement which was ap- parently interrupted on the high plains by rapid drying as the ice receded. A few Great Plains species did reach the moun- tains and are now found as isolated popu- lations (Fig. lb). Two genera of salamanders reached New Mexico and are found in the Jemez and Sacramento mountains of the southern part of that state: the genera are represented by Aneides hardyi and Pletho- don neomexicanus. One may speculate as to why these or other genera from this large family did not reach Colorado. The best an- swer available is that the warming at the close of the Ice Age provided moist, warm climatic conditions in what is now the southern Great Plains but less favorable conditions on the central plains. With the ice receding from the mountains, climatic conditions in the southern Great Plains of New Mexico were apparently more moist than at present and similar to the situation 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 45 in southern Nevada about 10,000 to 15,000 years ago. We are aware of only two plethodontid salamanders that have survived. The wide- spread tiger salamander may have been here during the Pleistocene. If it did exist here during the Pleistocene, it may repre- sent one of the few species able to survive that period by remaining in the region. Once the valleys became dry, the salaman- ders were isolated in the mountaintops with no opportunity to expand their range. Pre- sumably, their isolated mountain distribution and the rapidly drying conditions prevented them from reaching Colorado. To the west, north, and south of Utah the deserts devel- oped rapidly. Glacial lakes which were present in many Great Basin valleys dis- appeared or were reduced to salty rem- nants; associated vegetation changes isolated amphibians in small areas around waterways and desert springs (Figs. 2c, 2d, and 9). The drying out and warming of the southern portions of this vast inland area Fig. 2. Distribution of some amphibians in the southwestern United States: (a) Canyon treefrog, Hyhi arenicolor; (b) Woodhouse's toad, Bufo woodhoasei; (c) red-spotted toad, Bufo punctatus; and (d) south- western toad, Bufo microscaphus. 46 GREAT BASIN NATURALIST MEMOIRS No. 2 provided an opportunity for species in the southern deserts (northern Mexico and per- haps some areas in Sonora, Chihuahua, and Coahuila) to expand their ranges northward. It is now possible to detect some such range expansions along valleys running north and west from the international border. Time does not permit an examination of all species, but we can examine one. The leopard lizard, Crotaphytus wislizeni, ap- pears to have emerged from a refugium in the Chihuahua-Coahuila area and followed routes approximately as indicated by the ar- rows in Figure 3a. The migration resulted Fig. 3. (a) The theorized flow distribution for the leopard lizard Crotaphytus wislizeni; (b) geographical distribution as known today. Fig. 4. Present-day distribution of southwestern rep- tiles: (a) desert iguana, Dipsosaurus dorsalis; (b) desert tortoise, Gopherus agassizi, (c) zebratailed lizard, Calli- saurus draconoides; and (d) banded gecko, Coleonyx variegatus. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 47 in the distributional pattern for the species shown in Figure 3b. In this and other spe- cies occupation of some areas produced semi-isolation, and geographical subspecies have evolved in such areas as the Colorado Plateau, Virgin River Valley, and the Great Basin of Utah and Nevada. This is true not only for the leopard lizards but for most of the species that have extended their ranges into the northern and western valleys. Not all species moved as far or perhaps as fast. An examination of present-day ranges are the best indicators of the general movements that occurred. The following list of 23 species (Figs. 2 and 4-8) all show range extensions into the Great Basin and the valleys of the Colorado drainage. Some species have either expanded their range more rapidly or have been able to cross ele- vation barriers of 5,000 feet or higher (Figs. 5 and 6) while others have not (Figs. 4 and 8). Fig. 5. Distribution of southwest reptiles: (a) Desert spiny lizard, Sreloporus magister; (b) tree lizard, Urosaurus ornata; (c) side-blotched lizard, Uta stansburiana; and (d) less earless lizard, Holbrookia macu- 48 GREAT BASIN NATURALIST MEMOIRS No. 2 1. Southwestern toad (Bufo microscaphus) 2. Woodhouse toad {Bufo woodhousei) 3. Red-spotted toad (Bufo punctatus) 4. Canyon treefrog (Hyla arenicolor) 5. Desert tortoise (Gopherus agassizi) 6. Banded gecko (Coleonyx variegatus) 7. Desert iguana (Dipsosaurus dorsalis) 8. Zebra-tailed lizard (Callisaurus draconoides) 9. Lesser Earless lizard (Holbrookia maculata) 10. Desert spiny lizard (Sceloporus magister) 11. Side-blotched lizard (Uta stansburiana) 12. Tree lizard (Urosaurus ornata) 13. Western Whiptail (Cnemidophorus tigris) 14. Western Blind Snake (Leptotyphhps humilis) 15. Western Patch-nosed Snake (Salvadora hexdepis) 16. Coachwhip Snake (Masticophis flagellum) 17. Glossy Snake (Arizona elegans) 18. Common Kingsnake (Lampropeltis getulus) 19. Black-necked Garter Snake (Thamnophis cyrtopsis) 20. Western Ground Snake (Sonora semianulata) 21. Mojave Rattlesnake (Crotalus scutulatus) 22. Speckled Rattlesnake (Crotalus mitchelli) 23. Sidewinder (Crotalus cerastes) Some species must have survived the Pleistocene in refugia that lay between the Coachwhip WESTERN STATES Fig. 6. Distribution of southwestern reptiles: (a) Western patchnosed snake, Salvadora hexalcpis; (b) western blind snake, Leptotyphhps humilis; (c) coachwhip, Masticophis flagellum; and (d) western whip- tail, Cnemidophorus tigris. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 49 cold climates of the mountains and the drier, mild climates of southern plains. Such species include: 1. The Western Toad (Bufo boreas) 2. Spotted Frog (Rana pretiosa) 3. Rubber Boa (Charina bottae) 4. Western Garter Snake (Thamnophis Such species seemingly have moved north as climatic conditions permitted. They oc- cupied only mountain valleys in the south- ern parts of their present range (Fig. 9). Other species have apparently moved into the Intermountain Region from the northern or central Great Plains. Such spe- cies include: 1. The Chorus Frog (Pseudacris tnseriata) 2. Smooth Green Snake (Opheodrys vernalis) 3. Racer (Coluber constrictor) These are eastern species which seem to have entered through the northern Great Plains (Fig. 10). Presumably, the smooth green snake had a much wider distribution in earlier times than at present. This is in- dicated by its disjunct distribution. If there are reptile species that survived the Pleistocene in the lower valleys of the Great Basin of Utah and Nevada, the short- Fig. 7. Distribution of southwestern reptiles: (a) Glossy snake Arizona elegans; (b) black-necked garter snake, Tluimnophis cyrtopsis; (c) western ground snake, Sonora semianulata; and (d) common kingsnake, Lampropeltis getulus. 50 GREAT BASIN NATURALIST MEMOIRS No. 2 horned lizard (Phrynosoma dougkissi), the western skink (Eumeces skiltonianus), Fig. 10a), and the sagebrush lizard (Sceloporus graciosus) are the species most likely to have done so. These species now range from the valleys up to at least 9,000 feet in the mountains and plateaus. In summary, we can conclude that the ancestral stocks of the great majority of present day intermountain amphibians and Mojave Rattlesnake STERN STATES s \i* f^. 'V \ _ L I Sidewinder WESTERN STATE! ■Aj Speckled \ N Rattlesnake WESTERI Fig. 8. Distribution of southwestern reptiles: (a) Mo- jave rattlesnake, Crotalus scutulatus; (b) sidewinder, Crotalus cerastes; and (e) speckled rattlesnake, Crotalus mitchelli. reptiles originated either to the south or east of the area in question. By far the greater numbers of both amphibians and reptiles came from the south or the south- east. The Intermountain West is a good, if not a classical, example of the David Starr Jor- dan theory. He stated that animals have three alternatives if radical changes occur in the environment: 1. They can follow the environment and thus remain constant. 2. They can remain and adapt to the new environ- ment. 3. If they can do neither, they will become extinct. It may not be possible to cite an example of a reptile or amphibian that has remained constant. Yet we do have some that have wide distributions and little morphological divergence. Two examples are the Charina bottae and the Bufo boreas. Both have wide distribution with little external variation. Without a fossil record, we do not know how many amphibians and reptiles existed in our area since late Pleistocene time and were unsuccessful in the struggle for survi- val. Presumably, we have had during the last ten thousand to fifteen thousand years substantial environmental changes that of- fered challenges beyond the ability of some species to adapt. Other species have extend- ed their range through adaptive radiation, which increased the number of geographical subspecies or morphological clines in the species as isolated habitats were occupied. There is reason to believe that the move- ment north is still occurring. The estab- lishment of populations of Crotalus mitchelli and C. scutulatus in Utah appear to be re- cent (Fig. 8). The first specimen of C. scutulatus was taken in 1954 and the first C. mitchelli in 1960. Both were taken on the southwest slope of the Beaver Dam Mountains, only a few miles inside Utah. Since then, these species have apparently expanded their ranges and are seen more of- ten by field workers. The northern plateau Lizard (Sceloporus 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 51 u. elongatus) has in recent times crossed the occurs along the foothills extending north to central plateaus of Utah through the low Ephraim and south to Monroe. We do not areas between Emery and Salina and now have records of this species west of the Se- Western Toad ,'l STATES Fig. 9. Amphibians and reptiles with a more northern distribution but with populations extending south into parts of the Intermountain West and the Great Basin: (a) Spotted frog, Rana pretiosa; (b) rubber boa, Charina bottae; (c) western toad, Bufo boreas; and (d) common garter snake, Ttiamnophis sirtalis. 52 GREAT BASIN NATURALIST MEMOIRS No. 2 ■ELL 1978 INTEKMOL .STAIN BIOGEOGRAPHY! A SYMI'OSH M 53 vier River. A specimen of the long-nosed snake ( Rhniocheilus lecontet) was recentl) taken south of Dragerton- an indication that this species is still expanding its range. In conclusion, it should be noted that many areas in Utah souk- local, others ex- tensive) have had their reptile populations reduced by human activity. The most com- mon disruptive influence has been over- grazing on some private, Bureau of Land Management, and state lands. Figures lb, 3b, and 4-10 are taken largely from Stebbins 1966). Figure fa is from Etheridge fr961j. Even though the distribu- tion maps have not been brought up to date for 10 years, the ranges of species used in this study have changed little. LlTERAl i HI. Cll ED Ballinceb, R. K., and D. VV. Tinkle 1072. Systematica and evolution of tin- genus Ufa Sauna [guanidae] Misc Publ Mu* Zool Univ. Michigan 1 i~ I 83 Etheridce R L961 Late Cenozou glass lizards OphisaUTUi from the fOUtheni Great Plains. Herpetologica IT J T'j 186. I K R and W. W. Tanneb L975 Evolution of the- scleroporine lizards [guanidae ' Jreal Basin Nat 35 1 20 Stebbins R C L966 A held guide to western rep- tiles and amphibians Houghton Mifflin Co Boston Wells B. v., and C. D. Jobcensek 1964 Pleistocene wood rat middens and < limatic changes in the Moja ) i re ord of juni- per woodlands. Science 143 1171-1173. Fig. 10. Amphibians and reptiles which bav< irthem distribution hut which seemingly have entered the interrnouritain and Great Basin areas from the central Great Bla. mec&s skiltonumu.% this distribution is more comparable to those in Fig. 9 ; b smooth green snake, ftyh- eodrys i.errui': Cohtber "jnArutor: and d chorus frog. Pseudocrii mgfita. AVIAN BIOGEOGRAPHY OF THE GREAT BASIN AND INTERMOUNTAIN REGION William H. Behle' Abstract.— There are no endemic species of birds in the Great Basin. Nevertheless, a distinctive Great Basin avifauna exists which contains components of the Mojave Desert, Rocky Mountain, and Great Plains avifaunas as well as species obligate to sagebrush and the pinyon-juniper forest. Seemingly there has been little spread of Cal- ifornia and Sierra Nevada species eastward, but a westward extension from the Rocky Mountains of several spe- cies is indicated. While several Rocky Mountain species reach their western limits on the eastern edge of the Great Basin, others have extended into the eastern portion. Two Great Plains representatives are late arrivals, namely the Baltimore Oriole and Indigo Bunting, with evidence of introgression now occurring with related western species. A similar but longstanding situation exists for the flickers. A zone of hybridization occurs in northern Utah between two species of junco. A rather abrupt junction zone between the Great Basin and Mojave Desert avifaunas exists in southern Nevada and extreme southwestern Utah. Several species representing the Mo- jave Desert avifauna have extended their ranges in recent years into southern Utah. Geographically variable birds show diverse patterns of distribution along with much clinal variation and intergradation. A center of differen- tiation for four species occurs in western Utah in the eastern portion of the Great Basin while two more occur in the western portion of the basin. The Wasatch Front is a dividing area between western and eastern races in several species. Extreme southwestern Utah constitutes a transition area where several species are represented by different races or intergradational populations. A study of the avifaunas of 14 boreal "islands" in isolated moun- tain ranges in western and southeastern Utah in comparison with the Rocky Mountain "continent" in central and northern Utah shows a close correlation between number of species present and habitat diversity. In addition, a low correlation exists between the number of species that are permanent residents on isolated mountains and the distance of those mountains from the "continent." Biogeography is concerned with the dis- tribution of organisms in time and space. Applying this to birds and the Great Basin region, it is the consensus among students of avian paleontology that most species of modern birds arose during the Pleistocene (Selander 1965), but there is virtually no fossil record of birds for the Great Basin during that interval of time. Trimble and Carr (1961) mention that bones of birds as well as of several kinds of mammals and molluscs have been found in gravel over- lying the Raft Formation of American Falls Lake bed in southern Idaho which represent the late Quaternary, but no identities of the birds are given. A number of bird bones as- sociated with prehistoric human habitations in caves in the Great Basin have been found, two of the best-known sites being Danger Cave near Wendover (Jennings 'Department of Biology. University of Utah, Salt Lake City, Utah 84112. 1957) and Hogup Cave near the north- western corner of Great Salt Lake (Aikens 1970), but all the bird bones and feathers represent living species. Hall (1940) de- scribes an ancient nesting site of White Pelicans at Rattlesnake Hill on the north- eastern edge of the town of Fallon, Nevada, containing bones of White Pelicans, Double- crested Cormorants, and a Canada Goose. The bones were situated beneath a water- formed calcareous layer, which indicated that the bones had been under water at least once; but whether this was before, at, or after the time when Lake Lahonton at- tained its maximum level was not ascer- tained. The implication from the find is that the avian associates in this prehistoric time were about the same as one finds today at the colony on Anaho Island in nearby Pyra- mid Lake. Despite the virtual absence of a 55 56 GREAT BASIN NATURALIST MEMOIRS No. 2 fossil record, it is probably safe to assume that the species of birds present in the Great Basin in the Quaternary were essen- tially the same as those present in the re- gion today. With different climatic condi- tions, however, from time to time there have doubtless been different assemblages of birds and different distributional patterns than are seen at present. Thus, in the ab- sence of a fossil record for the region under consideration, reliance must be placed on an analysis of the distribution of today's species in the search for clues to dispersal routes and subspecific differentiation. Before considering the spatial dimension of the biogeography of birds of the region under consideration, it may be well to note two special items in connection with birds. One is that some birds are migratory. Thus a distinction must be made between sum- mer residents and permanent residents. The migratory summer residents are able to eas- ily traverse distances between mountain ranges and so are less subject to the effects of isolation than are the sedentary per- manent residents. The second point is that there is a wealth of data pertaining to the distribution of birds in the collections of many museums, with much of the data readily available in published reports. For the region under consideration the following constitute the principal sources of informa- tion on the distribution of birds: for Califor- nia, Grinnell and Miller (1944); for Nevada, Linsdale (1936 and 1951) and Johnson (1965, 1973, 1974); for Idaho, Burleigh (1972); for Utah, Behle (1943, 1955, 1958, 1960), Behle, Bushman and Greenhalgh (1958), Behle and Ghiselin (1958), Behle and Perry (1975), and Hay ward, Cottam, Wood- bury, and Frost (1976); for Colorado, Bailey and Niedrach (1965). Phillips (1958) has dis- cussed many special problems having to do with the collecting of birds and the short- comings of museum collections. Even though the material available falls short of the need, birds are still one of the best known groups of animals in terms of biogeography. Great Basin Avifauna Turning now to the spatial dimension, an important initial consideration is whether there is a distinctive avifauna in the Great Basin. Are the kinds of birds that occur in western Utah and Nevada different en masse from those found in the California- Sierra Nevada region on the west or the Colorado-Rocky Mountain region to the east? This query pertains only to land birds, since water birds are widespread in their occurrence throughout North America and generally show few regional distinctions ex- cept for relative abundance of particular species. An analysis of the distribution of the land birds indicates that the great ma- jority that occur in the Great Basin range widely throughout western North America. There are about 154 kinds of resident birds in this wide-ranging category. Any dis- tinctions, then, pertain to a relatively few kinds, but mostly it is a matter of a differ- ent combination of species in the Great Ba- sin as compared with the assemblages in neighboring regions. Udvardy (1963: 1157) includes a Great Basin avifauna in his treat- ment of the bird faunas of North America, stating that the species fall geographically and ecologically into two groups, namely (1) the sagebrush-arid woodland faunal group and (2) the northwestern arid wood- land faunal group. Miller (1951) in his anal- ysis of the distribution of the birds of Cali- fornia includes a Great Basin avifauna as one of four faunal groups represented in the state, one that is intrusive into northeastern California east of the Sierran crest. He states that the Great Basin avifauna consists of two categories: (1) species of interior continental derivation that occur south of or below the boreal areas, and (2) geographic- races that have differentiated in the Great Basin at austral levels. He designates 35 kinds as belonging to the Great Basin avi- fauna. Johnson (1975) followed Miller in his treatment of a Great Basin avifauna. Probably the most distinctive feature about the Great Basin avifauna is the pres- 1978 INTERMOUNTAIN BIOGEOGRAPHYI A SYMPOSIUM 57 ence of certain birds that are associated with two plant formations that occur widely throughout the region, namely big sage (Ar- temisia tridentata) and the pinyon-juniper woodland. Birds that occur almost exclu- sively in stands of sagebrush are the Sage Grouse, Sage Thrasher, and Sage Sparrow. Birds that occur chiefly, if not exclusively, in the pygmy woodland, which itself has much sage interspersed with the junipers and pinyon pines, are the Cassin's Kingbird, Gray Flycatcher, Scrub Jay, Pinyon Jay, Plain Titmouse, Bush-tit, Blue-Gray Gnat- catcher, Cedar Waxwing, Gray Vireo, Black-throated Gray Warbler, and Brewer's Sparrow. However, the pygmy woodland occurs throughout the Southwest so these associated species of birds occur in areas beyond the Great Basin. To properly char- acterize the Great Basin avifauna, com- parisons with surrounding regions are neces- sary. There are about 30 kinds of distinctive birds that occur in the California-Pacific Coast-Sierra Nevada region that are not known to occur in either the Great Basin or the Rocky Mountains. Many are endemic to the West Coast area and constitute the most conspicuous elements of the California avifauna. Some of these, such as the Moun- tain Quail and White-headed Woodpecker, occur in the Sierra Nevada on the western rim of the Great Basin, but I find little evi- dence of these distinctive California forms spilling over eastward into the mountain ranges in the Great Basin. Several northern birds reach the southern limits of their ranges, at least in part in the Great Basin. These are the Marsh Hawk, Roughed Grouse, Sharp-tailed Grouse, Sage Grouse, Lewis Woodpecker, Tree Swallow, Swain- son's Thrush, Water Pipit, American Red- start, and Fox Sparrow. Many southern birds reach their northern limits, in at least part of their range, in the Great Basin. These are the Whip-poor-will, Black Phoebe, Gray Flycatcher, Plain Titmouse, Bewick's Wren, Bendire's Thrasher, Blue- gray Gnatcatcher, Gray Vireo, Virginia's Warbler, Black-throated Gray Warbler, Painted Redstart, Scott's Oriole, Lesser Goldfinch, Black-throated Sparrow, Gray- headed Junco, and Black-chinned Sparrow. There are about 25 kinds representing the Mojave Desert avifauna that occur in south- eastern Nevada and southwestern Utah but which do not penetrate any farther north into the Great Basin except on an acciden- tal basis. These are discussed in the follow- ing section of this paper. Nineteen kinds of birds occur in Colorado that do not occur as breeders in either the Great Basin or California areas. Mostly these are species of the Great Plains avi- fauna that reach the western limits of their ranges along the east base of the Rocky Mountains. One species is endemic to the mountains of Colorado, namely the Gray- crowned Rosy Finch. There are several spe- cies that occur as breeders in both the Rocky Mountains and the Great Basin which do not occur in the California-Sierra Nevada area. Thus they reach their western limits within the Great Basin. These are the Northern Three-toed Woodpecker, Catbird, Brown Thrasher, Veery, Water Pipit, Black Rosy Finch and Indigo Bunting. None of these are common in the Great Basin and at least two, the Brown Thrasher and Indigo Bunting, appear to be late arrivals in the region west of the Rocky Mountains. Three species are found at the eastern edge of the Great Basin in Utah but are not known to occur in the basin per se. These are the Purple Martin, Gray Jay, and Pine Gros- beak. Several kinds are essentially restricted in Utah in their breeding range to the Colo- rado River drainage system, but occasionally individuals occur in the Great Basin as acci- dentals. These are the Gambel Quail, Costa Hummingbird, Roadrunner, Bendire's Thrasher, and Blue Grosbeak. Finally, I know of no species of bird that is endemic to the Great Basin. From all this we can conclude that there is a distinctive Great Basin avifauna but it is one that is not characterized by endemic- species. Rather it is recognizable on the 58 GREAT BASIN NATURALIST MEMOIRS No. 2 basis of a different assemblage of birds, many of which are intrusive from surround- ing regions. There is more evidence of a western spread of eastern species into the Great Basin than there is of an eastward spread from the Sierra Nevada-California area. Because of the lack of endemics, the Great Basin avifauna is not as distinctive as surrounding avifaunas, but it is more sharp- ly confined, being delimited on the west by the Cascade-Sierra cordillera and on the east by such outlying ranges of the Rockies as the Wasatch Mountains of northern Utah and the high plateaus of central Utah. On the south the Great Basin avifauna meets the Mojave Desert avifauna in a rather dis- tinct and narrow junction zone. There is no comparable junction zone or mountain bar- rier at the northern limits of the Great Ba- sin. Here the Great Basin species gradually merge with those of either the western woodland edge or those of the open Palouse country east of the Cascades. Relations of Mojave Desert and Great Basin Avifaunas in Southwestern Utah and Southeastern Nevada The northern limits of the Mojave Desert Biome in Nevada have been mapped by Gullion et al. (1959: 279). Areas included are Meadow Valley Wash, Muddy River, and Pahranagat Valley. In southwestern Utah, the warm southern desert occurs along the floor of the Virgin River Valley to the mouth of Zion Canyon near Spring- dale (including Coal Pits Wash) as well as along the lower stretches of tributary streams such as La Verkin, Ash, and Santa Clara creeks and Beaver Dam Wash on the west side of the Beaver Dam Mountains. In Arizona it occurs along the Virgin Riv- er Valley. There are 28 kinds of summer resident birds in this region that are repre- sentatives of the Mojave Desert avifauna. Fifteen of these are known to occur in Utah only in this area. The other 13 occur there regularly but a few extralimital records exist elsewhere in the state. The Mojave Desert avian indicators are the Black Hawk, Gam- bel's Quail, White-winged Dove, Ground Dove, Inca Dove, Roadrunner, Lesser Nighthawk, Costa's Hummingbird, Rivoli's Hummingbird, Ladderbacked Woodpecker, Wied's Crested Flycatcher, Black Phoebe, Vermilion Flycatcher, Verdin, Cactus Wren, Le Conte's Thrasher, Crissal Thrasher, Black-tailed Gnatcatcher, Phainopepla, Bell's Vireo, Lucy's Warbler, Painted Red- start, Hooded Oriole, Scott's Oriole, Sum- mer Tanager, Blue Grosbeak, Abert's Tow- hee, and Rufous-crowned Sparrow. Several of these species seem to have extended their ranges into southwestern Utah in recent years, namely, the Black Hawk, White- winged Dove, Inca Dove, Rivoli's Hum- mingbird, Wied's Crested Flycatcher, Black- tailed Gnatcatcher, Summer Tanager, and possibly the Rufous-crowned Sparrow, al- though the latter may represent an over- looked species associated with a relict grass- land habitat. A summary of records and details of distribution for this complement of birds has recently been presented by Behle (1976b) for the three-state region. Im- mediately to the north of this Mojave Desert or Lower Sonoran area and at high- er elevations in the region in the pinyon- juniper belt, birds are found that represent the Great Basin avifauna. There are some aspects of subspecies dis- tribution and intergradation in extreme southwestern Utah that are significant in terms of southern derivations of the popu- lation. These are discussed elsewhere in this paper. The hybridization that produces in- tergradation in these several species, as well as increased variability in the populations, suggests the presence of a suture zone, us- ing the terminology of Remington (1968). He defined a suture zone as "a band, whether narrow or broad, of geographic- overlap between major biotic assemblages, including some pairs of species or semi- species which hybridize in the zone." As Uzzell and Ashmole (1970) further note, su- ture zones stand to biotas as zones of sec- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM ondary intergradation stand to pairs of pop- ulations. Unless one prefers to regard the Gilded Flicker as a separate species from the Red-shafted Flicker, to my knowledge no hybridization occurs in extreme south- western Utah at the species level. Rather, the crossing is between representatives of different subspecies producing intermediate and highly variable populations where, in addition to the intergrades, typical repre- sentatives of the two parental stocks occur. In the region to the north of the Virgin River Valley, some cases are known where introgression has taken place. These are dis- cussed in another section of this paper. Boreal Islands and Effects of Isolation One of the most significant aspects of zoogeography in the Great Basin and Inter- mountain Region pertains to the dis- continuous occurrence of boreal species on the many isolated mountaintops of the re- gion. The distribution of birds on 31 such islands has been discussed by Johnson (1975) in a study patterned after similar studies by Brown (1971) on mammals of the Great Ba- sin ranges and Vuilleumier (1970) on birds in the paramo islands in the northern Andes. Although Johnson had data available from several of my reports for certain is- lands in western Utah which constituted the eastern fringe of his study area, additional data for Utah have been mobilized for this paper to extend Johnson's study. Although I have followed his procedures, our data are not precisely comparable because of region- al differences in the avifauna, my elimi- nation of water birds from the boreal cate- gory, and the circumstance that I have followed Brown's approach of considering as boreal species those that occur above 7500 feet elevation rather than attempting to determine the lower edge of the forest woodland. The 80 species that I have desig- nated as boreal are listed in the Appendix along with an indication of their presence or absence on the 14 boreal islands studied in western and southeastern Utah. The basic- data for the several islands are presented in Table 1. These data were first subjected to a normality check which showed that they fit a normal distribution in an untrans- formed condition. The data were then ana- lyzed by means of a partial correlation analysis which showed that three variables, namely elevation of highest peak, total area, and habitat diversity score (HDS) were highly correlated. Then a stepwise multiple regression study showed that HDS had the highest correlation with the total number of bird species occurring. The R-value (correla- tion coefficient) was .86, which was signifi- cant at the .001 level. Because of the inter- correlation among the three independent variables, the multiple regression analysis was run first with HDS included in the equation while excluding elevation and total area. Then it was run excluding HDS but including elevation and area. Finally calcu- lations were made including all three varia- bles. As is indicated by the data summa- rized in Table 3 and the adjusted R2 values, the effect of multicolinearity is present when all three variables are included in the equation. Although not as strong, these re- sults follow those of Johnson closely. Be- cause of the multicolinearity when all three independent variables were included in the equation, the R2 value given in Figure 2 (which is .75) is the value derived from the equation excluding elevation and total area. The results shown in Figure 2 are sim- ilar to those of Johnson (1975: 553). Although there is a high correlation be- tween the number of kinds of birds occur- ring and general habitat diversity as repre- sented by the habitat diversity scores, there is the complication that the HDS involves many environmental variables. In attempt- ing to identify particular aspects of the hab- itat community structure that control the kinds and number of species present, John- son (1975: 555) analyzed the species compo- sition of the boreal birds and their ecologic rolls in the community. He divided them into two groups: "Restricted," which oc- curred in 5 or fewer of his 31 sample areas, 60 GREAT BASIN NATURALIST MEMOIRS No. 2 Fig. 1. Map of Utah showing locations of Boreal Islands and Rocky Mountain Continent area above 7500 feet elevation. 1. Raft River Mts., 2. Deep Creek Mts., 3. Stansbury Mts., 4. Oquirrh Mts., 5. House Range, 6. Needle Range, 7. Wah Wah Mts., 8. Frisco Mts., 9. Mineral Mts., 10. Pine Valley Mts., 11. La Sal Mts., 12. Abajo Mts.-Elk Ridge, 13. Henry Mts., 14. Navajo Mtn., 15. Wasatch-Uinta-Tushar-High Plateau Continent. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 61 and "Standard," which occurred in 28 or more. Birds in the "Standard" category are presumed to have generalized boreal re- quirements in contrast to specialized re- quirements for the "Restricted" group. Johnson noted Willson's (1974) work that deals with aspects of habitat structure in re- lation to species and numbers of birds. A more recent paper along similar lines is Flack's (1976) study of bird populations in the aspen forests in western North America. The approach of Willson and Flack focuses attention on the significance of particular environmental variables presently covered by Johnson's habitat diversity score. The next highest correlation shown by my data is with width of barrier, but this is significant only in connection with the cate- gory of permanent residents (R = .43). In other words, for the summer residents there is no correlation between number of kinds occurring on an island and distance from the nearest island or continent, while for the permanent residents the number of kinds decreases with remoteness from the continental area. The distance correlation is minor, however, compared with that for habitat diversity. Again my results are es- sentially the same as those of Johnson (1975). He expressed the opinion that the distance factor in the case of birds operates through impoverishment of habitat rather than through ease of access. A low correlation shows up for my data between number of species and total area of the island (see Table 2). This is contrary to the results of both Brown and Johnson as well as the postulate of MacArthur and Wilson (1963, 1967) that area and environ- mental diversity are closely related and that total area serves as a good general predictor of habitat variety. The lack of correlation between number of species and size of area for the islands that I studied was probably influenced by the disparate results for the two smallest islands, namely the Frisco Mountains with a size of 11 square miles and only 19 kinds of birds as compared to Navajo Mountain with 13 square miles and 49 kinds of birds. I gave a habitat diversity score of 3 to the Frisco Mountain area and a 5 to Navajo Mountain. The Frisco range is very dry and has a sparse coniferous for- est. Navajo Mountain is also lacking in sur- face accumulation of water, yet supports much more forest covering. Environmental Table 1. Data for Boreal Islands and the Rockv Mountain Mainland in Utah." Area Mountain No. Range N, N2 N3 AR WB DM EHP LHP HDS 1 Raft River Mts. 61 22 39 64 48 79 9892 41.92 10 2 Deep Creek Mts. 52 20 32 223 9 104 12101 39.83 11 3 Stansbury Mts. 44 18. 26 54 16 39 11031 40.27 8 4 Oquirrh Mts. 50 20 30 82 16 19 10676 40.22 9 5 House Range 25 9 16 25 35 63 9725 39.09 4 6 Needle Range 29 10 19 92 11 65 9783 38.16 4 7 Wah Wah Mts. 42 15 27 54 11 53 9065 38.33 6 8 Frisco Mts. 19 8 11 11 11 .38 9669 38.31 3 9 Mineral Mts. 34 12 22 24 25 11 9619 38.20 5 10 Pine Valley Mts. 46 18 28 79 39 10 10325 37.32 10 11 La Sal Mts. 64 25 39 314 28 42 13089 38.26 12 12 Abajo Mts.— Elk Ridge 42 22 20 368 38 70 11445 37.50 9 13 Henry Mts. 41 17 24 108 38 21 11615 38.07 7 14 Navajo Mtn. 49 22 27 13 62 58 10416 37.02 5 15 Wasatch-Uinta Mainland 80 33 47 - - - 13498 40.77 18 °N, = total number boreal species found; N2 = number of these permanently resident; N3 = no. summer residents; AR = total area above 7500 feet in square miles; WB = width of interisland lowland desert barrier, e.g., distance from closest boreal island; DM = distance from mainland; EHP = elevation of highest peak; LHP = latitude of highest peak; HDS = Habitat Diversity Score. GREAT BASIN NATURALIST MEMOIRS No. 2 patchiness or some other aspect of the more extensive woodland on Navajo Mountain presumably accounts for the greatly in- creased number of species present. In these two instances, at least, total area is not as good a predictor of number of kinds of birds as is total forest woodland area with all the attendant attributes, whatever they may be. From his study of boreal mammals on the mountaintops of the Great Basin ranges, Brown (1971) concluded that their diversity and distribution could not be explained in terms of an equilibrium between coloniza- Table 2. Results of partial correlation analysis of island data. Upper number indicates the correlation coefficient; lower number is the level of significance. Meaning of symbols is the same as in Table 1. N, N2 N3 AR WB DM EHP LPH HDS 1.000 N, .001 .933 1.000 N2 .001 .970 .001 .816 1.000 N3 .001 .442 .001 .581 .001 .313 1.000 AR .114 .029 .275 .001 .319 .428 .220 -.023 1.000 WB .267 .127 .449 .938 .001 .167 .169 .153 .329 -.018 1.000 DM .569 .564 .602 .251 .952 .001 .581 .673 .474 .771 .055 .128 1.000 EHP .029 .008 .087 .001 .851 .662 .001 .329 .145 .430 -.132 -.204 .305 -.016 1.000 LHP .250 .622 .125 .653 .485 .289 .956 .001 .867 .832 .824 .655 .091 .138 .714 .323 1.000 HDS .001 .001 .001 .011 .757 .638 .004 .260 .001 Table 3. Summary of results from stepwise multiple regression analysis showing relationship of total number of bird species (the dependent variable) to independent variables. HDS = habitat diversity score; WB = width of interisland barrier; EHP = elevation of highest peak; AR = total area; DM = distance from mainland; LHP = latitude of highest peak. Variable Multiple R R2 R2 Change Simple r Treatment A. EHP and AR excluded as independent variables HDS .86693 .75156 .75156 .86693 WB .89979 .80963 .05807 .31877 LHP .90712 .82286 .01323 .32918 (constant) Treatment B. HDS exc luded as an independent variable AR .44173 .19512 .19512 .44173 LHP .58987 .34795 .15283 .32918 WB .72392 .52405 .17611 .31877 DM .74207 .55067 .02661 .16689 (constant) Treatment C. All variables included in the analysis HDS .86693 .75156 .75156 .86693 WB .89979 .80963 .05807 .31877 AR .91085 .82964 .02002 .44173 DM .91693 .84076 .01111 .16689 (constant) .91870 .84402 .00326 .58097 EHP 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 63 tion and extinction. His interpretation was that boreal mammals reached all the islands during the Pleistocene and since then there have been extinctions but no colonizations. In his study of boreal birds, Johnson (1975) concluded that a similar nonequilibrium sit- uation prevails for the permanent resident species, but for the summer residents the equilibrium theory of island species number does apply since species are excluded by habitat deficiencies rather than barriers. Subspecies of Geographically Variable Species in Utah An aspect of biogeography that is of pri- mary interest to the systematist is the geo- graphic distribution of different subspecies or races of geographically variable species. Twenty-three species present systematic- problems in Utah. Of these, 7 are montane or boreal forms, 14 are valley or austral species, and 2 are wide-ranging types that extend from the valleys up to the mountain- tops. Of the total, 14 are represented by 2 breeding races and 4 by 3 races, with possi- bly another in the last category. Another 4 species are represented by only one race in the state, but each has an intergrading pop- ulation in some part of Utah that is transi- tional with another race in surrounding re- gions. The distribution of the races and populations in nine geographic regions in the state is indicated in Table 4 except for the Red Crossbill, about which a decision as to the number of races represented in Utah Habitat Diversity Score Fig. 2. Relationship between habitat diversity score (HDS) and total number of birds (N\) occurring on boreal islands in Utah. Numbers of sample areas correspond to those used in Fig. 1 and Table 1. R2 value shown is the adjusted value because of the small number of sample areas. 64 GREAT BASIN NATURALIST MEMOIRS No. 2 awaits the results of a pending systematic review by Allen Phillips. Areas of inter- gradation of varying extent occur between the races. Two instances of a minor barrier effect have been revealed. No uniform pat- tern of distribution prevails. Rather, there are several situations indicated whereby 2 or more species show racial changes in about the same general area. The picture of variation is more indicative of broad changes on a regional basis than of differen- tiation in isolated mountain ranges, as is of- ten the case with more sedentary groups such as mammals. One distributional pattern is where differ- ences occur between populations in the west desert portion of northern Utah and those of the Wasatch and Uinta mountains. This is seen in the Dusky Grouse, Cliff Swallow, Mountain Chickadee, Brown Cree- per, Scrub Jay, and Steller's Jay. Cliff Swal- lows represent an extreme case of gradual clinal variation, with only specimens from the ends of the cline in extreme western Tahi.k 4: Subspecies of geographically variable birds or intermediate populations represented in various geography House, Pine Valley Raft River, Stansbury, Needle, Mountains Deep Creek Oquirrh Wah Wah Vircin River Valley Mountains Mountains Mountains Beaver Dam Wash Ruteo jamaicensis calurus calurus calurus calurus Red-tailed Hawk Dendragapus ohscurus oreinus oreinus > - obscurus Blue Grouse obscurus Ottts aslo inyoensis inyoensis inyoensis inyoensis > Screech Owl yumanensis Bubo virginianus occidentalis occidentalis occidentalis patlescens Great Homed Owl Chordeiles minor hesperis hesperis hesperis henryi Common Nighthawk Picoides vilbsus leucothorectis > monticola leucothorectis leucothorectis Hairy Woodpecker monticola Empidonax traillii adastus adastus adastus > extimus Willow (Traill's) extimus Flycatcher Eremophila alpestris utahensis utahensis utahensis - Homed Lark Petrochelidon pyrrhonota hypopolia hypopolia > hypopolia > tachina Cliff Swallow pyrrhonota pyrrhonota Cyanocitta stelleri - macrolopha macrolopha macrolopha Steller's Jay Aphelocoma cocrulescens nevadae nevadae nevadae nevadae Scrub Jay Parus atricapillus nevadensis nevadensis nevadensis nevadensis Black-capped Chickadee Parus gambeli inyoensis inyoensis > inyoensis > inyoensis Mountain Chickadee wasatchensis wasatchensis Certhia familiaris leucosticta leucosticta leucosticta leucosticta Brown Creeper Catherpes mexicanus - - - conspersus Canyon Wren Suriiu mexicana _ - - - Western Bluebird / anius ludovicianus gambeli gambeli gambeli gambeli Loggerhead Shrike (leothlypis trichas occidentalis occidentalis occidentalis occidentalis > Common Yellowthroat scirpicola Agelaius phoeniceus fortis > fortis fortis fortis > Red-winged Blackbird nevadensis nevadensis Molothnis aler artemisiae artemisiae artemisiae > obscurus Brown-headed Cowbird obscurus Ctirpixlacus mexicanus solitudinus solitudinus solitudinus solitudinus House Finch Melospiza meUxiiu montanus montanus montanus fallax Song Sparrow 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 65 and eastern Utah sufficiently different to be assigned to separate races (see Behle 1976a). In two species more of a step cline is repre- sented. In one of these, the Dusky Grouse, specimens from the Deep Creek Mountains near the Utah-Nevada border, are typical of the race Dendragapus obscurus oreinus. Those from the Oquirrh Mountains are clos- est to oreinus but show an approach to ob- scurus. In the Wasatch Mountains the grouse represent the race obscurus. A sim- ilar situation exists in the Mountain Chick- adee. Those from the Deep Creek Moun- tains represent the race Varus gambeli inyoensis. Those from the Stansbury and Oquirrh mountains are closest to inyoensis but show an approach to wasatchensis which occurs in the Wasatch Mountains and thence east to the Uinta Mountains. In the Brown Creeper, representatives from all the west desert ranges represent the race Cer- thia familiaris leucosticta. Those from the Wasatch Mountains are a highly variable lot of intergrades but as a whole stand closest Wasatch Mountains Wasatch Plateau Pavant, Tushab Mountains Aquarius, PaUNSAUGUNT, Mabkacunt Plateaus I'inta Mountains— Tavaputs Plateau La Sal— Ahajo Henhv Mountains calurus calurus calurus fuertesi " obscurus obscurus obscurus obscurus obscurus inyoensis inyoensis inyoensis inyoensis inyoensis occidentalis occidentalis occidentalis occidentalis pallescens hesperis hesperis henryi howelli henryi monticola monticola monticola > leucothorectis monticola luecothorectis adastus adastus > extimus adastus > extimus adastus adastus > extimus utahensis leucolaema leucolaema leucolaema leucolaema occidentalis hypopolia > pyrrhonota macrolopha > annectens woodhouseii > nevadae hypopolia > pyrrhonota macrolopha woodhouseii > nevadae hypopolia > pyrrhonota macrolopha woodhouseii > nevadae pyrrhonota > hypopolia macrolopha woodhouseii hypopolia > pyrrhonota macrolopha woodhouseii nevadensis nevadensis nevadensis garrinus garrinus wasatchensis wasatchensis wasatchensis wasatchensis gambeli montana > leucosticta montana montana montana montana conspersus > griscus conspersus conspersus bairdii conspersus occidentalis conspersus bairdii gambeli gambeli gambeli gambeli > excubitorides gambeli > excubitorides occidentalis occidentalis occidentalis occidentalis occidentalis fortis fortis fortis fortis fortis artemisiae artemisiae artemisiae artemisiae artemisiae solitudinus > frontalis solitudinus > frontalis solitudinus > frontalis solitudinus > frontalis frontalis montanus montanus montanus montanus montanus GREAT BASIN NATURALIST MEMOIRS No. 2 to the race montana. The Steller's Jays of the west desert ranges including the Oquirrh Mountains are typical of the race Cyanocitta stelleri macrolopha, while those from the Wasatch Mountains constitute an intergrading population between mac- roloplia and annectens, a northern race. The Scrub Jays of the Oquirrh Mountains and other west desert ranges are typical of the race Aphelecora caerulescens nevadae, but those from the Wasatch are intergrades be- tween nevadae and woodhouseii, closest to the latter. Thus in these several species a break occurs along the west escarpment of the Wasatch Mountains dividing west desert races from intergrading populations in the Wasatch Mountains and eastward. The Jor- dan Valley between the Oquirrh and Wasatch Mountains, only about 25 miles across, thus seems to act as a weak barrier for the montane forms. The second pattern is for the break along a west-east cline to occur farther east be- tween the Wasatch and Uinta mountains. Here there is not even a valley to serve as the line of demarcation. This situation is seen in the Red-tailed Hawk and Black- capped Chickadee. For the hawk, the popu- lation in the Wasatch and all of western Utah represents the race Buteo jamaicensis calurus, while those from the Uinta Basin and Tavaputs Plateau region are closest to fuertesi, a race which extends southeast into Texas. The Black-capped Chickadee of the Oquirrh and Wasatch ranges represents the race Parus atricapillus nevadensis. By the time the Uinta Basin is reached the popu- lation represents garrinus. The third pattern is for a race or popu- lation to be represented in northern Utah and a different one in the southern part of the state. Exemplifying this are the Steller's Jay, Hairy Woodpecker, and Great Horned Owl. As previously noted, the Steller's Jays from the Wasatch Mountains represent an intergrading population between the races annectens and macrolopha, closest to the latter. In southern Idaho the jays are closest to annectens. South of Mount Nebo at the southern end of the Wasatch Mountains, the jays are typical of macrolopha. In the case of the Hairy Woodpecker the break occurs south of the Aquarius, Paunsaugunt, and Markagunt plateaus. This is farther south than the transition area for the Steller's Jays. An unexpected distributional feature of the Hairy Woodpecker is that the southern race Picoides villosus leucothorectis ex- tends farther north in the isolated mountain ranges of the Great Basin in western Utah than it does in the plateaus and mountains of central Utah. This suggests that the prop- agules for the west desert ranges came from the southeast rather than directly west from the Wasatch, a situation similar to that of the Three-toed Woodpecker (Johnson 1975: 548). Whatever the direction of spread, leu- cothorectis and monticola seem to have met in the Snake and Deep Creek ranges where an intergrading population occurs. In con- trast, a sharp break in the distribution of the two subspecies occurs between the Tushar and Mineral mountains in south- western Utah. The population of the Tushar Mountains is monticola while that of the Mineral Range about 25 miles to the west with Beaver Valley between is typical leu- cothorectis. In the case of the Horned Owls, the race extending across southern Utah is Bubo virginianus pallescens but its range swings north in eastern Utah to include the La Sal Mountain-Moab region. The fourth situation is found in extreme southwestern Utah along the Virgin River Valley, where, in addition to the numerous indicator species of the Mojave Desert avi- fauna previously discussed, there are differ- ences at the subspecies level for several kinds of birds. In three geographically vari- able species there are races that do not oc- cur elsewhere in the state. These are a sub- species of Cliff Swallow (Petrochelidon pyrrhonota tachina), a race of Brown Cow- bird (Molothrus ater obscurus), and a race of Song sparrow Zonotrichia melodia fallax). These three races are of southern origin. In four other species, the populations are inter- gradational toward southern races. This is 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 67 the case for the Screech Owl, where the population is Otus asio inyoensis toward yumanensis. The Gilded Flicker (formerly Colaptes chrysoides but now considered to be C. aaratus chrysoides) has been observed in Beaver Dam Wash in Utah, and one specimen has been obtained that is inter- mediate between that race and the Red- shafted Flicker (C. a. cafer). In the Rough- winged Swallow, the population is Stelgi- dopteryx ruficoUis serripennis toward psam- mochroa. In the Yellow- throat, the popu- lation is Geothlypis trichas occidentalis toward scirpicola. Yet another intergrading population occurs in the region in the case of the Red-winged Blackbird, but the race with which intergradation occurs is a west- ern race. The population is Agelaius phoe- niceus fortis toward nevadensis. A fifth distributional pattern shows a dif- ferent population in the Red Rock country of southeastern Utah as compared to the rest of the state. This is seen in the Night- hawks and Horned Larks. For the former, the race in southeastern Utah is Chordeiles minor henryi as opposed to howelli to the north in the Uinta Basin and hesperis in western Utah. For the Horned Larks there is an intergrading population between the race Eremophila alpestris leucolaema and occidentalis in southeastern Utah that is closest to leucolaema. Finally, a situation occurs in the Horned Larks that is unlike the racial distribution of any other geographically variable species in the state. One race, leucolaema, occurs in subalpine meadows in the plateaus of cen- tral Utah and in alpine tundra of the Tush- ar Mountains, while a different race, uta- hensis, occupies the desert floor of the valleys below. The high elevation race is the same as the lowland race of the Uinta Basin in northeastern Utah. This same phe- nomenon of two races at different altitudes in the same general region is found in the Sierra Nevada (see Behle 1942), where the race sierrae occurs in montane meadows as opposed to different races in the lowland vallevs both east and west of the mountains. In contrast, in the Raft River Mountains of northwestern Utah, Horned Larks taken from the top of the mountain at 9500 feet represent the same race as in the lowlands, namely utahensis; and, in the Colorado Rockies, the race leucolaema ranges from the valleys up to the Arctic tundra at over 11,000 feet. Clinal Variation Clines are essentially a phenomenon of geographic variation but they are also part of the picture of biogeography inasmuch as they would not be evident if samples were not present from many geographic areas. Clinal variation is manifest in the characters of many kinds of birds in the Great Basin and Utah. Occasionally similar clines appear in unrelated species, which suggests that some common environmental influence is exerting a selective influence. Clines in some instances extend in a north-south di- rection, while in others they extend from east to west or northeast to southwest. Phil- lips (1958) mentions the Song Sparrow (Zono- trichia melodia) in the Great Basin as an example where two clines cross per- pendicularly. One cline toward longer wings and darker color extends northward while another toward short wings, large bill, and heavy breast-spotting proceeds westward. In connection with his work on the birds of Nevada, Linsdale (1938: 175) itemized the changes observed for several variable spe- cies, then generalized that for many birds there is a decrease in size toward the south. The largest individuals occur in the north- eastern corner of the state. The bill be- comes shorter and stubbier toward the east and smaller toward the south. The wings and tail are generally longer toward the east. General coloration becomes paler and grayer toward the east and sometimes brighter and darker in the vicinity of the Colorado River. In Utah, clines are most evident in size characters. The usual pattern is for birds in the northern part of the state to be of 68 GREAT BASIN NATURALIST MEMOIRS No. 2 larger size than those in the southern part, with a smooth gradient occurring the length of the state. The gradient in Utah is usually a portion of a more extensive cline extend- ing throughout western North America. A recent study that I made of the White- throated Swift (Behle 1973) revealed clinal variation nicely. Measurements of popu- lations from Montana south to Arizona were analyzed. Clinal variation was most appar- ent in wing length, which is regarded in or- nithological systematics as a good indicator of overall size. Clinal variation was less evi- dent in tail length and virtually nonexistent in bill and tarsal lengths. For wing length, the means for the several populations mea- sured showed a gradual transition from 143.2 mm in the Montana sample to 136.5 in the Arizona-New Mexico sample, a dif- ference of 6.7 mm. While there was a gen- eral decrease in wing length from north to south in Utah samples, a mosaic pattern of variation was shown in the several semi-iso- lated populations represented. For example specimens from the Raft River Mountains in northwestern Utah have the longest wings in the state (average wing length 146.0 mm). They are larger than those from cen- tral northern Utah, northeastern Utah, or Colorado and are closest to the Montana population in size. Swifts from the Beaver Dam Wash in extreme southwestern Utah have the shortest wing length (wing 134.2). They are smaller than samples from central southern and southeastern Utah and are even smaller than the Arizona-New Mexico sample. These extreme Utah populations differ in average wing length by 11.8 mm, which is greater than that between Mon- tana and Arizona-New Mexico birds (6.7 mm). The circumstance that northern swifts have longer wings than do southern swifts may be correlated with the behavioral fea- ture that northern individuals migrate dur- ing the winter from their breeding areas while those in the southern part of their range are sedentary. Another case of north- south clinal variation in size is seen in the Cliff Swallows in western North America (Behle 1976a). Clinal variation in size in Utah has become apparent from our studies of the Great Horned Owls and Hairy Woodpeckers (unpublished data). Clines are also evident in Utah birds in color characters. A west-east gradient oc- curs in several species in northern Utah whereby paler-colored birds occur in the desert Great Basin portion of the state, with a transition eastward to darker birds in the Wasatch and Uinta mountains. Such clines are most evident in dorsal coloration. Spe- cies showing this phenomenon are the Dus- ky Grouse, Screech Owl, Common Night- hawk, Cliff Swallow, Horned Lark, Scrub Jay, Mountain Chickadee, and Creeper. Of the lot, the phenomenon seems to be most pronounced in the Dusky Grouse. Represen- tatives are pale and gray in the ranges of eastern Nevada and in the Deep Creek Mountains of western Utah. In the Oquirrh Mountains they start to be slightly darker, showing more brown. The darkening is ac- centuated in the Wasatch Mountains and continues to a still greater degree in the Uinta Mountains and eastward into Colo- rado. In general, east-west clinal variation in Utah is the reverse of that for Nevada, since the birds become paler and grayer in the western part of the state, where the Great Basin occurs. Clinal variation in color from darker birds in the north to lighter birds in the south shows up in a few birds such as the Steller's Jay. In Utah, as in Ne- vada, brighter coloration occurs in some species in the valley of the Virgin River. Examples showing this are Yellow-throats and Song Sparrows. Secondary Contact of Species IN THE INTERMOUNTAIN REGION In recent years several studies have been made of secondary contact of pairs of close- ly related species or subspecies of birds in North America (see Selander 1965: 536, for a listing of kinds and sources of informa- tion). For most cases, the area of contact is the Great Plains, but in four instances the 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM phenomenon shows up in Utah. Three of these involve eastern and western kinds hybridizing. The fourth involves a northern and a southern species meeting. For two of the four the contact has probably been brought about during the past few decades; the other two are of longer duration. The first case pertains to the Indigo Bunting (Passerina cyanea, a species that originally occurred only in eastern North America), and the Lazuli Bunting (P. amoena), a west- ern species. Apparently the planting of trees and shrubs in cities and parks across the plains states bridged the former grasslands hiatus separating the two species, and a highway was thus provided for dispersal of the Indigo Bunting westward into the range of the Lazuli Bunting. Historical records suggest that the Indigo Bunting arrived in Utah about 40 years ago. That hybridization of the two species has occurred is indicated by two intermediate specimens. One, in the Cornell University collection, was taken near Ogden, Weber County, Utah, on 12 August 1945 (Sibley and Short 1959: 447). The other is in the University of Utah col- lection and was taken along Minnie Maud Creek, 2 miles east of Nutter's Banch Duchesne County, Utah, on 30 June 1966. The Indigo Bunting is now fairly common in southern Utah, where it exists sympa- trically with the Lazuli Bunting. Whitmore (1975) has recently discussed the inter- specific behavioral competition now in evi- dence in this region between the two spe- cies. The second instance of recently estab- lished contact in Utah pertains to two kinds of oriole, the Baltimore Oriole, formerly called Icterus galbula, which is an eastern type, and the Bullock's Oriole, formerly des- ignated as I. bullockii, a common western kind. Worthen (1973) reported an example of the Baltimore Oriole taken 2 miles south of Milford, Beaver County, Utah, on 27 June 1964. It was one of a series of several orioles obtained at this location. Although in worn plumage, the specimen represents a "pure" first-year male. While this particular specimen shows no tendency toward the Bullock's Oriole, some others in the series do show evidences of hybridization. Sibley and Short (1964) have shown that hybridiza- tion in the two orioles is now common throughout the Great Plains. As a result, the two orioles are presently considered as races of one species, e.g., I. g. galbula and 7. g. bullockii. The third case of hybridization in Utah pertains to flickers. There are three types of flickers in North America: the Yellow- shafted Flicker, essentially an eastern bird formerly designated as Colaptes auratus; the Bed-shafted Flicker of the west, formerly called C. cafer; and the Gilded Flicker of the southwest and lower California, for- merly called C. chrysoides. The Yellow- shafted and Bed-shafted forms for over 100 years have been known to hybridize in a broad montane belt in western North Amer- ica extending from British Columbia south- ward throughout the Bocky Mountain re- gion. Short (1965) interprets the picture of speciation as follows. He postulates a geog- raphic separation of the ancestral auratus- cafer population during the Illinois glacial age or earlier. The separation continued during subsequent periods of glaciation (ex- cept for possible hybridization between the two differentiated stocks during interglacial periods). With the waning of the last major advance of the Wisconsin period of glacia- tion, the eastern Yellow-shafted Flicker, auratus, was able to expand its range west- ward and northwestward into British Co- lumbia. In contrast, the Bed-shafted Flicker, cafer, remained restricted to the area south of the glaciers in the western United States. Eventually the two populations made con- tact and hybridized along the length of the Bockies. Utah is west and south of the main zone of introgression and Short scarcely mentions the area in his discussion, but there is evidence of much crossing taking place in Utah. A recent study by Bich (1967) of 137 specimens in the University of Utah collection revealed that 85 are "pure" cafer, 4 are "pure" auratus, and 48 are in- 70 GREAT BASIN NATURALIST MEMOIRS No. 2 termediates. This is a surprisingly large number of intergrades with so few auratus seemingly present. It suggests that there is- little or no selective pressure against the characters produced by auratus genes. The greatest flow of auratus genes into the cafer population pool in Utah appears to be oc- curring in northwestern Utah, as indicated by the highest incidence of intermediates. In contrast, there are fewer intermediates from eastern Utah, suggesting that the east- west gradient from the Great Plains area is not of great significance in Utah. In other words, the Yellow-shafted Flickers in Utah have seemingly come mostly from the northwest rather than the east. Inter- mediates occur throughout the Great Basin mountain ranges. Short (1965) conceives of all the North American flickers belonging to a single spe- cies, Colaptes auratus, and, following his lead, the Yellow-shafted Flicker is now known as C. a. auratus, the Red-shafted Flicker is C. a. cafer, and the Gilded Flick- er is C. a. chrysoides. The latter apparently hybridizes with the Red-shafted Flicker in extreme northwestern Arizona and south- western Utah; Wauer and Russell (1967) re- port a specimen taken at the Terry Ranch in Beaver Dam Wash, Washington County, on 28 April 1965 as being a hybrid of chry- soides X cafer derivation. The last case of secondary contact in Utah involves two species of junco that come together in extreme northern Utah, with a relatively restricted zone of hybridi- zation extending in an east-west direction across the state. One population is a north- ern form, a race of the Dark-eyed (Oregon) Junco (Junco hyemalis mearnsi). The other is a southern form, the Gray-headed Junco (/. caniceps caniceps). The contact of these two kinds was originally detected by Miller (1941: 200). At a locality 10 miles east of Kamas, all examples that he collected were pure caniceps. There was a shift in frequen- cy of characters of /. h. mearnsi northward indicated by specimens from 20 miles north of Kamas, in the Uinta Mountains, then in succession Woodruff, Randolph, and Garden City, until nearly pure populations of mearnsi were found near the Utah- Idaho border. Since then breeding hybrid juncos have been taken in the Wasatch Mountains east of Salt Lake City and in the Uinta Mountains. Hybrids extend westward in northern Utah to the Raft River Mountains and beyond into northeastern Nevada. Centers of Differentiation in the Great Basin On a previous occasion (Behle 1963), I studied the distribution of the races of 50 geographically variable species of birds whose ranges include or impinge upon the Great Basin and its flanking regions. The re- sults showed that the Great Basin is not in itself one large center of differentiation. In- stead, several distribution areas were re- vealed that either occur in portions of the Great Basin or are situated in nearby sur- rounding regions. The areas were designated as the Warner Region, Sierra Nevada, Western Great Basin, Eastern Great Basin, Rocky Mountains, Northern Idaho, Inyo Re- gion, Mojave Desert, Colorado Desert, and Navajo Country. The races of the geograph- ically variable species occurring in each of the 10 distribution zones were listed. Since many species occupied each area, it was the different combinations of races along with common transition zones or areas of inter- gradation between races that served to characterize and delimit the several areas. No one species showed conformance in the distribution of its races with the various dis- tribution areas outlined, although the horned lark approached this in slight de- gree. In only a few instances were races en- demic in any one area. The Great Basin is too diversified in terms of environmental factors to have influenced in some common way all the geographically variable birds that are found in the region. The differen- tiation and distribution of the races is pre- sumably correlated largely with localized different environments, but, in addi- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 71 tion, barrier effects and individual histories of the various kinds of birds in terms of their point of origin, dispersal, and depen- dency on particular plant associations have played a role. There are indications of three centers of differentiation in the Great Basin region, one in the eastern part, and two in the western portion. The latter two have been evaluated by Miller (1941). One is the White Mountain area of eastern California in the southwestern portion of the Great Basin. The other is the Warner Mountain- Warner Valley region of southern Oregon and northeastern California in the north- western portion of the Great Basin. More recently Johnson (1970) presented additional data for the avifauna of the Warner Moun- tains and has reevaluated the affinities of the boreal elements. He gives different re- sults than I presented (Behle 1963). The center of differentiation in the eastern part of the Great Basin rests on four races that have fairly common, though not identical, ranges. Three of these were described in the course of our fieldwork at the University of Utah, namely a race of Dusky Grouse (Den- dragapus obscurus oreinus), a race of Horn- ed Lark (Eremophila alpestris utahensis), and a race of Fox Sparrow (Passerella iliaca swarthi). The fourth is a race of Black- capped Chickadee (Parus atricapillus neva- densis) described by Linsdale of the Univer- sity of California. Theoretical Historical Aspects of Distribution of Birds in the Great Basin To my knowledge, no direct evidence has been detected of the influence of Pleisto- cene or Holocene climates on the distribu- tion of birds in the Great Basin or Inter- mountain Region. Still, some inferences can be drawn. Resident birds present in the re- gion today are closely associated in their occurrence with particular biotic commu- nities, especially the plant components. Since climatic change has resulted in altera- tion of community types, the avian associ- ates have almost certainly been affected too, either being forced out of areas where the plant habitat has disappeared or in- vading new areas as their requisite habitat has become established. Dispersal resulting from population pressure has also resulted in extensions of ranges. In some instances former allopatric species have become sym- patric, as in the cases of the flickers and buntings. With contractions of ranges, for- mer sympatric species conceivably have be- come allopatric. Such movements would be in the nature of long-term adjustments. Of particular interest in this connection is the present-day discontinuous distribution of the coniferous forest on the mountaintops of the Great Basin and the attendant boreal species of birds. Johnson (1975: 556) has noted the two theories that have been of- fered to account for this. One proposes that during the Pleistocene cold climates pre- vailed with relatively more moisture and less evaporation than in the region today. These conditions induced the formation of glaciers in the mountains and the pluvial lakes Bonneville, Lahonton, and a host of lesser lakes on the floor of the Great Basin. At the same time, the coniferous forest pre- sumably extended altitudinally down into the valleys, bordering on the lakes, and be- came distributed more or less continuously throughout the Great Basin. Boreal species of birds presumably accompanied the con- iferous forest and also occurred more or less continuously at lowland elevations. Sub- sequent climatic change to the warmer and drier conditions of today resulted in melting of the glaciers, disappearance or diminution of the lakes, and retreat of the coniferous forest up into the mountains. The boreal birds presumably were also forced to move up into the mountains, where they occur as breeders today. Martin and Mehringer (1965) have mobilized the evidence from pollen studies in support of such climatic changes. In accord with this line of reason- ing such avian species as the northern Three-toed Woodpecker, Water Pipit, and 72 GREAT BASIN NATURALIST MEMOIRS No. 2 Black Rosy Finch were presumably formerly much more widespread and abundant but have subsequently been confined to the mountaintops where Hudsonian zone or al- pine-arctic conditions prevail. Concurrently the lowland valleys were invaded by low- land species from surrounding regions, spe- cies adapted to the warmer, xeric conditions that have come to prevail there. The second point of view is that the montane pockets of boreal forest and at- tendant faunas have come about through dispersal over long distances from parental stocks such as those in the Sierra Nevada and Rocky mountains. Some evidence per- taining to the vegetation to support this in- terpretation has been presented by Wells and Berger (1967) and Critchfield and Al- lenbaugh (1967), and Johnson (1975) has noted the probable role of certain species of boreal birds such as the Band-tailed Pigeon, Pinyon Jay, and Clark's Nutcracker in long- distance colonization through transporting and/or burying seeds of conifers. In accord- ance with this theory, the Three-toed Woodpecker, Water Pipit, and Black Rosy Finch have extended their ranges westward from the Rocky Mountain continental area only to certain boreal islands in the eastern part of the Great Basin. Present indications are that the Rosy Finch has progressed the farthest, being known from the Jarbidge Mountains, the Ruby Mountains, and the Wheeler Peak area of the Snake Range in eastern Nevada. The Water Pipit stops at the Deep Creek Range in extreme western Utah. The Northern Three-toed Wood- pecker is known from the Snake Range. However, I suspect that if more collecting were done, all three species would be found at Wheeler Peak and the Ruby Mountains in Nevada. I favor the relict mountaintop theory as Brown (1971) does for mammals. As Johnson notes, these two hypotheses are not mutually exclusive. Both processes could have occurred in the past. Extensions of range are occurring at present, as indicated by the historic record for certain species. The diversity of distribution patterns that have been noted for Utah leads to the infer- ence that each species has had its own par- ticular distributional history determined by its habitat requirements and the habitat changes experienced. M Imi ANAGEMENT IMPLICATIONS Birds come into the picture of natural re- source management in the Great Basin in several ways. In connection with environ- mental impact studies, special consideration is being given to threatened and endangered species such as the Bald Eagle, Peregrine Falcon, and Osprey. All of these occur in the Great Basin. Indeed, studies of raptors by Brigham Young University biologists have revealed high populations for many species in remote areas of the Great Basin. Even subspecies are important in terms of endangered species, because it is the south- ern race of Bald Eagle and the southern race of Peregrine Falcon that are endan- gered. Another area where subspecies are important in management pertains to the introduction of exotic game species. If more should be introduced (and there are serious objections to the practice) stock should be selected representing races whose native habitat is most nearly like that where the introduction is to be made. Another point is that populations at type localities, such as the Dusky Grouse (which is a game species) from the Deep Creek Mountains, should be protected. Regional avifaunistic reports, such technical papers as descriptions of new forms and systematic revisions, and sym- posia such as the present one, especially the published proceedings, constitute valuable resource material for those charged with making evaluations. They provide baseline data to compare against in future years to establish long-term changes. In the 130 years since settlement of the Great Basin and other parts of the Inter- mountain West, many well known changes have occurred in the vegetation as pointed out by Cottam (1947) and others. Con- comitant changes have occurred in the bird 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 73 life. For example, as the grassland was es- sentially extirpated from Utah through over- utilization, the habitat for the Grasshopper Sparrow and Sharp-tailed Grouse was re- moved, the result being the near extermina- tion of these kinds of birds in the region today. As pertains to the grouse, certain protected areas containing the little remain- ing requisite habitat are needed for its sur- vival. One such tract is east of Wellsville, Cache County, Utah. As more and more sagebrush is removed for land cultivation, Sage Grouse and other sage-inhabiting spe- cies are being affected. The Conservation Committee of the Wilson Ornithological So- ciety (see Braun et al. 1976) has recently re- ported on the extent of alteration of this community and attendant deleterious effects on the associated bird life. Chaining out of junipers and pinyon pine in the Great Basin is of less consequence in terms of the bird life because the extent of the habitat is so vast. Nevertheless, I feel that some typical areas should remain undisturbed to serve as study areas. They should be large enough to preserve habitat diversity and maintain spa- tial relations intact. Yet extensive areas of continuous forest may not be as effective in preserving communities and species as would numerous, smaller, diversified, irregu- lar areas. A large, essentially undisturbed and diversified area is the Wheeler Peak re- gion in the Snake Range in eastern Nevada. At one time the area was proposed for a Great Basin natural park. I would like to see this area so designated. This would af- ford some measure of protection. The Leh- man Cave National Monument has already been established there. The Bureau of Land Management is, I believe, considering the designation of the Deep Creek Mountain in western Utah as a quasi-primitive area. A complication is that part of that range is In- dian reservation. The Beaver Dam Wash area of extreme southwestern Utah is unique in its flora and fauna and needs pro- tection—especially from collectors. Regarding such theoretical considerations as the application of island biogeography theory to conservation practice as has been advocated in the design of wildlife refuges, Simberloff and Abele (1976) express the opinion that the application of the general principle is premature at the present time. They feel that, in this particular instance, broad generalizations have been based on limited and insufficiently validated theory and on field studies of taxa which may be idiosyncratic. The implication is that much more research is needed. I suggest that the boreal islands of the Great Basin constitute propitious areas for further avian research as a sequel to Johnson's and my work. Summary and Conclusions Although there are no endemic species of birds in the Great Basin region of western North America, nevertheless a distinctive avifauna exists there by virtue of a different combination of birds as well as the presence of many species associated with sagebrush and the pinyon-juniper forest. Physiographic boundaries determine the limits of the Great Basin avifauna on the west and east, while on the south there is a sharp junction zone with the Mojave Desert avifauna that occurs in southern Nevada and in extreme southwestern Utah. To the north there is a gradual blending with the avifauna of the i Palouse prairie and the northern montane woodland. About 30 kinds of distinctive birds that occur in the California-Pacific Coast-Sierra Nevada region are not known to occur in the Great Basin, suggesting that relatively little eastward spread has oc- curred. In contrast, seven Rocky Mountain species reach their western limits within the Great Basin. Some of the latter group, namely the Yellow-shafted Flicker, Balti- more Oriole, and Indigo Bunting, are recent arrivals and introgression has occurred with western congeners. Instability of present-day ranges for many species of birds is further indicated by the finding in recent years of several other kinds, mostly in southwestern Utah, that are new to the state list. Ten northern species reach their southern 74 GREAT BASIN NATURALIST MEMOIRS No. 2 limits in at least part of their ranges in the Great Basin. A zone of hybridization be- tween a race of the northern Dark-eyed Junco (/. h. mearnsi) and the southern Gray- headed Junco (/. c. caniceps) occurs across northern Utah and northeastern Nevada. Sixteen southern species reach their north- ern limits in the Great Basin, while an addi- tional 25 species stop at a distinct Great Basin— Mojave Desert junction zone in southern Nevada and extreme southwestern Utah. Three avifaunas are represented in the Great Basin region today, namely the Bocky Mountain, Great Basin, and Mojave Desert. Montane species, which are mostly associ- ated with the coniferous forest, are dis- continuously distributed in boreal islands on the tops of isolated mountain ranges in the general region. An analysis of the avifaunas of 14 such islands in western and south- eastern Utah, as compared with that of the Bocky Mountain continent in central and northern Utah, shows a close correlation be- tween the number of species present and habitat diversity. A slight, negative correla- tion shows up for permanent residents with distance from the continent. The results are similar to those of Johnson (1975) for a dif- ferent set of islands mostly located in Ne- vada. An analysis of the distribution of races of 22 species in Utah represented by more than one race in the state reveals a variety of patterns. For several a break occurs along the Wasatch Front on the east side of the Great Basin between a west desert race and either an eastern montane race or an intergrading population toward a different race in eastern Utah. In a few others, the break is farther east between the Wasatch and Uinta mountains. Another situation is for there to be one race in northern Utah and a different race in the southern part of the state. In three species, there are differ- ent races or populations in southeastern Utah; but southwestern Utah is the most distinctive transitional area where, in three species, different races are represented and in five others intergradational populations occur. For the Horned Lark, one race oc- curs in subalpine meadows in central Utah, and a different race is a summer resident in the desert region at the base of the moun- tains. In some species intergrading popu- lations occur over broad areas; in others the phenomenon is confined to a narrow zone. A center of differentiation occurs in west- ern Utah in the eastern portion of the Great Basin where four races of geographically variable birds have ranges that somewhat coincide. This is similar to the White Mountain and Warner Mountain centers in the western portion of the Great Basin. Clinal variation occurs in many species, in- volving both size and color characters. Some clines run north and south and others run east and west. Some are gradual; others are step clines. Past climatic change has doubtless influenced the distribution of spe- cies and avifaunas in the region. It is infer- red that during cold intervals of the Qua- ternary boreal birds occurred in lowland valleys, but with a warming trend they have retreated to the mountaintops, where they are found today. This would account for the current distribution of the Water Pi- pit and Black Bosy Finch, although the pos- sibility exists of a westward spread of these species from the Bocky Mountains. Acknowledgments I am indebted to many people for help in various ways in the preparation of this pa- per. Tom Boner and Marie Magleby com- piled the lists of species for the various bo- real islands. John Wyckoff mobilized the data for physical features in Table 1, han- dled the statistical treatment of the data, and constructed the graph for Figure 2. Dave Prouse and Marie Magleby worked on the map. Magleby and William Pingree measured hundreds of birds, and Pingree helped in making subspecies determinations. Norma Fernley did the typing. Many indi- viduals have helped in the field work 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 75 throughout the years, team effort. It has indeed been a Literature Cited Aikens, C. M. 1970. Hogup Cave. Univ. Utah Anthro- pol. Pap. No. 93: 1-286. Bailey, A. M., and R. J. Niedrach. 1965. Birds of Colorado, Vol. 1, 2. Denver Museum of Natural History, Denver. Behle, W. H. 1942. Distribution and variation of the horned larks (Otocoris alpestris) of western North America. Univ. Calif. Publ. Zool. 46: 205-316. 1943. Birds of Pine Valley Mountain region, southwestern Utah. Univ. Utah Biol. Ser. 7 (5): 1-85. 1955. The birds of the Deep Creek Moun- tains of central western Utah. Univ. Utah Biol. Ser. 11 (4): 1-34. 1958. The birds of the Raft River Mountains, northwestern Utah. Univ. Utah Biol. Ser. 11 (6): 1-40. 1960. The birds of southeastern Utah. Univ. Utah Biol. Ser. 12 (1): 1-56. 1963. Avifaunistic analysis of the Great Basin region of North America. Proc. 13th Intern. Or- nithol. Congr.: 1168-1181. 1973. Clinal variation in White-throated Swifts from Utah and the Rocky Mountain re- gion. Auk 90: 299-306. 1976a. Systematic review, intergradation, and clinal variation in Cliff Swallows. Auk 93: 66-77. 1976b. Mohave Desert avifauna in the Virgin River Valley of Utah, Nevada, and Arizona. Condor 78: 40-48. Behle, W. H., J. B. Bushman, and C. M. Greenhalgh. 1958. Birds of the Kanab area and adjacent high plateaus of southern Utah. Univ. Utah Biol. Ser. 11 (7): 1-92. Behle, W. H., and J. Ghiselin. 1958. Additional data on the birds of the Uinta Mountains and Basin of northeastern Utah. Great Basin Nat. 18:1-22. Behle, W. H., and M. L. Perry. 1975. Utah birds: Checklist, seasonal and ecological occurrence charts, and guides to bird finding. Utah Mu- seum of Natural History, Salt Lake City. Benson, S. B. 1935. A biological reconnaissance of Navajo Mountain, Utah. Univ. Calif. Publ. Zool. 40: 439-456. Blanchard, J. F. Jr. 1973. An ecologic and system- atic study of the birds of the Oquirrh Moun- tains, Utah. Unpublished master's thesis, Uni- versity of Utah, Salt Lake City. Braun, C, M. F. Baker, R. L. Eng, J. S. Gashwiler, and M. H. Schroeder. 1976. Conservation Committee report on effects of alteration of sagebrush communities on the associated avi- fauna. Wilson Bull. 88: 165-171. Brown, J. H. 1971. Mammals on mountaintops: nonequilibrium insular biogeography. Amer. Naturalist 105: 467-478. Burleigh, T. D. 1972. Birds of Idaho. Caxton Print- ers, Ltd., Caldwell, Idaho. Critchfield, W. B., and G. L. Allenbaugh. 1969. The distribution of Pinaceae in and near northern Nevada. Madrono 19: 12-26. Cottam, W. P. 1947. Is Utah Sahara bound? Bull. Univ. Utah, Biol. Ser. 37 (11): 1-40. Flack, J. A. D. 1976. Bird populations of aspen for- ests in western North America. Amer. Ornithol. Union, Ornithological Monogr. 19: 1-97. Grinnell, J., and A. H. Miller. 1944. The distri- bution of the birds of California. Pacific Coast Avif. 27: 1-608. GULLION, G. W., W. M. PULICH, AND F. G. Evenden. 1959. Notes on the occurrence of birds in southern Nevada. Condor 61: 278-297. Hall, E. R. 1940. An ancient nesting site of the White Pelican in Nevada. Condor 42: 87-88. Hayward, C. L., C. Cottam, A. M. Woodbury, and H. H. Frost. 1976. Birds of Utah. Great Ba- sin Nat. Mem. 1: 1-229. Jennings, J. D. 1957. Danger Cave. Univ. Utah An- thropol. Pap. 27: 1-328. Johnson, N. K. 1965. The breeding avifaunas of the Sheep and Spring ranges in southern Nevada. Condor 67: 93-124. 1970. The affinities of the boreal avifauna of the Warner Mountains, California. Occas. Pap., Biol. Soc. Nevada 22: 1-11. 1973. The distribution of boreal avifaunas in southeastern Nevada. Occas. Pap., Biol. Soc. Nevada 36: 1-14. 1974. Montane avifaunas of southern Ne- vada: historical change in species composition. Condor 76: .334-337. 1975. Controls of number of bird species on montane islands in the Great Basin. Evolution 29: 545-563. Linsdale, J. M. 1936. The birds of Nevada. Pacific Coast Avif. 23: 1-145. 1938. Bird life in Nevada with reference to modifications in structure and behavior. Condor 40: 173-180. 1951. A list of the birds of Nevada. Condor 53: 228-249. MacArthur, R. H., and E. O. Wilson. 1963. An equilibrium theory of insular zoogeography. Evolution 17: 373-387. 1967. The theory of island biogeography. Princeton University Press, Princeton, N.J. 76 GREAT BASIN NATURALIST MEMOIRS No. 2 Martin, P. S., and P. J. Mehringer, Jr. 1965. Pleistocene pollen analysis and biogeo- graphy of the southwest, pp. 433-451. In: H. E. Wright, Jr. and D. G. Frey (eds.), The Qua- ternary of the United States. Princeton Univ. Press, N.J. Miller, A. H. 1941. Speciation in the avian genus Junco. Univ. Calif. Publ. Zool. 44: 173-434. 1951. An analysis of the distribution of the birds of California. Univ. Calif. Publ. Zool. 50: 531-644. Phillips, A. R. 1958. The nature of avian species. J. Arizona Acad. Sci. 1: 22-30. Remington, C. L. 1968. Suture-zones of hybrid in- teraction between recently joined biotas. Evolu- tionary Biol. 2: 321-428. Rich, W. J. 1967. The woodpeckers of Utah. Un- published master's thesis, University of Utah, Salt Lake City. Selander, R. K. 1965. Avian speciation in the Qua- ternary, pp. 527-542. In: H. E. Wright Jr. and D. G. Frey (eds.), The Quaternary of the United States. Princeton Univ. Press, N.J. Short, L. L., Jr. 1965. Hybridization in the flickers (Colaptes) of North America. Bull. Amer. Mus. Nat. Hist. 129: 307-428. Sibley, C. C, and L. L. Short, Jr. 1959. Hybridization in the buntings (Passerina) of the Great Plains. Auk 76: 443-463. 1964. Hybridization in the orioles of the Great Plains. Condor 66: 130-150. Simberloff, D. S., and L. G. Abele. 1976. Island biogeography theory and conservation practise. Science 191: 285-286. Trimble, D. E., and W. J. Carr. 1961. Late Qua- ternary history of the Snake River in the Ameri- can Falls region, Idaho. Bull. Geol. Soc. Amer. 72: 1739-1748. Udvardy, M. D. F. 1963. Bird faunas of North America. Proc. 13th Intern. Ornithol. Congf. 13: 1147-1167. Uzzell, T., and N. P. Ashmole. 1970. Suture- zones: an alternative view. Systematic Zool. 19: 197-199. Vuilleumier, F. 1970. Insular biogeography in con- tinental regions. I. The northern Andes of South America. Amer. Naturalist 104: 373-388. Wauer, R. H., and R. C. Russell. 1967. New and additional records of birds in the Virgin River Valley. Condor 69: 420-423. Wells, P. V., and R. Berger. 1967. Late Pleisto- cene history of coniferous woodland in the Mo- jave Desert. Science 155: 1640-1647. Whitmore, R. C. 1975. Indigo Buntings in Utah with special reference to interspecific com- petition with Lazuli Buntings. Condor 77: 509-510. Willson, M. F. 1974. Avian community organiza- tion and habitat structure. Ecology 55: 1017-1029. Woodbury, A. M., and H. N. Russell, Jr. 1945. Birds of the Navajo country. Univ. Utah Biol. Ser. 9 (1): 1-160. Worthen, G. L. 1968. The taxonomy and distribu- tion of the birds of the southeastern Great Ba- sin, Utah. Unpublished master's thesis, Univer- sity of Utah, Salt Lake City. 1973. First Utah record of the Baltimore Oriole. Auk 90: 677-678. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 77 Occurrence of boreal birds on montane islands and portion of Rocky Mountain Continent in Utah. 1 * i 1 I I I I I I J J 1 I i | 1 i S i I " S ^ - >° U I a "S ^ i a. cc Q Si 3 c b « ^ "C j: .S «-£ « m^ g "Goshawk (Accipiter gentilis) XX X XX XXXX "Sharp-shinned Hawk (Accipiter striatus) XXXXX X XXXXXX "Cooper's Hawk (Accipiter cooperii) XXXX X X XX XX "Red-tailed Hawk (Buteo jamaicensis) XXXXX XXXXXX XX "Golden Eagle (Aquila chrysaetos) XXXXX X XX XX "American Kestrel (Falco sparverius) XXXX XXXXX "Dusky Grouse (Dendragapus obscurus) XXXX XXXX X "Ruffed Grouse (Bonasa utnbellus) X X "Band-tailed Pigeon (Columba fasciata) , XXXXXX "Flammulated Owl (Otus flammeolus) X XXX "Great Horned Owl (Bubo virginianus) XXX XXX XX "Pygmy Owl (Glaucidium gnoma) X "Spotted Owl (Strix occidentalis) X X "Long-eared Owl (Asia otus) XX X "Saw-whet Owl (Aegolius acadicus) XXX X White-throated Swift (Aeronautes saxatalis) XX X XXXXXXXXX Species ° = Permanent residents. Others are summer residents. 78 GREAT BASIN NATURALIST MEMOIRS No. 2 Black-chinned Hummingbird (Archilochus alexandri) Broad-tailed Hummingbird (Selasphorus platycercus) Calliope Hummingbird (Stellula calliope) "Common Flicker (Colaptes auratus) "Pileated Woodpecker (Dryocopus pileatus) "Yellow-bellied Sapsucker (Sphyrapicus varius) "Williamson's Sapsucker (Sphyrapicus thyroideus) "Hairy Woodpecker (Picoides villosus) "Downy Woodpecker (Picoides pubescens) "Northern Three-toed Woodpecker (Picoides tridactylus) Hammond's Flycatcher (Empidonax hammondii) Dusky Flycatcher (Empidonax oberholseri) Western Flycatcher (Empidonax difficilis) Western Wood Peewee (Contopus sordidulus) Olive-sided Flycatcher (Contopus borealis) Horned Lark (Eremophila alpestris) Violet-green Swallow (Tachycineta thalassina) Tree Swallow (Tachycineta bicolor) Purple Martin (Progne subis) "Gray Jay (Perisoreus canadensis) "Steller's Jay (Cyanocitta stelleri) "Clark's Nutcracker (Nucifraga columbiana) "Black-capped Chickadee (Parus atricapillus) "Mountain Chickadee (Parus gambeli) "White-breasted Nuthatch (Sitta carolinensis) "Red-breasted Nuthatch (Sirta canadensis) "Pygmy Nuthatch (Strta pygmaea) "Brown Creeper (Certhia familiaris) House Wren (Troglodytes aedon) XX X XX XXXXX X XXX XXX X xxxxxxxxxxxxxxx X X XX xxxxxxx X X xxxx xxxxxxxxxx X X X X X X XXXX X XXX X XXXX X XXXXXXX X XXX X XXXX X X X XX XXXX xxxxxxx XX XX XXXX X XX xxxxxxxxxxxxxxx X X XX XX X XX X XXXX X xxxxxxx xxxxxxxx xxxxxx XXXX XX X xxxxxxxxxxxxxxx XX xxxxxx xxxxxxxxxxxxxxx X xxxxxx XXXX X XXXX XX xxxxxxx xxxxxxx 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 79 Rock Wren (Salpinctes obsoletus) American Robin (Turdus migratorius) Hermit Thrush (Catharus guttatus) Swainson's Thrush (Catharus ustulatus) Veery (Catharus fuscescens) Western Bluebird (Sialia mexicana) Mountain Bluebird (Sialia currucoides) Townsend's Solitaire (Myadestes townsendi) Golden-crowned Kinglet (Regulus satrapa) Ruby-crowned Kinglet (Regulus calendula) Water Pipit (Anthus spinoletta) Solitary Vireo (Vireo solitarius) Warling Vireo (Vireo gilvus) Orange-crowned Warbler (Vermivora celata) Virginia's Warbler (Vermivora virginiae) Yellow-rumped Warbler (Dendroica coronata) Grace's Warbler (Dendroica graciae) X X MacGillivray's Warbler (Geothlypis tolmiei) Wilson's Warbler (Wilsonia pusilla) Western Tanager (Piranga ludoviciana) Black-headed Grosbeak (Pheucticus melanocephalus) Cassin's Finch (Carpodacus cassinii) "Pine Grosbeak (Pinicola enucleator) Black Rosy Finch (Leucosticte atrata) "Pine Siskin (Carduelis pinus) "Red Crossbill (Loxia curvirostra) Green-tailed Towhee (Pipilo chlorura) Rufous-sided Towhee (Pipilo erythrophthalmus) Vesper Sparrow (Pooecetes gramineus) xxxxxxx xxxxxx xxxxxxxxxxxxxxx xxxxxxxxxxxxxxx XX XX X X X XX XXXXXXXXXXXXX X XX X XX XXXX X XXX X XX XXXXXXXXXXXX XX X XX XXXX X XXX XX XXXXXXX XXX XXX XXXX X X X XXXX XX XXX XXX xxxxxxxxxxxxxxx XXXX XX xxxxxx X XXXX XX xxxxxxx XXXX X xxxxxxx XXXXXXXX XXX XX X X X XX XXXXXXXXXXX XXX XXXX xxxxx XXXXXXX XXX X X XXX X XXX XXX XX XXXX 80 GREAT BASIN NATURALIST MEMOIRS No. 2 Dark-eyed Junco (Junco hyemalis) X Gray-headed Junco (Junco caniceps) XXXX XXXXXXXXXX Chipping Sparrow (Spizella passerina) XXXXXXXXXXXXXXX White-crowned Sparrow (Zonotrichia leucophrys) XXX X XX X Fox Sparrow (Zonotrichia iliaca) XX XX Lincoln's Sparrow (Zonotrichia lincolnii) X XX Song Sparrow (Zonotrichia melodia) XX XX X Totals 61 52 44 50 25 29 42 19 .34 46 64 42 41 49 80 THE FLORA OF GREAT BASIN MOUNTAIN RANGES: DIVERSITY, SOURCES, AND DISPERSAL ECOLOGY K. T. Harper', D. Carl Freeman1, W. Kent Ostler1, and Lionel G. Klikoff2 Abstract.— The high elevation floras of 9 mountainous "mainlands" (3 in the Sierra-Cascade system and 6 in the High Plateau-Wasatch-Teton system) and 15 isolated mountain "islands" in the Intermountain Region have been analyzed. Mainland floras support more species per unit area and show a smaller increase in diversity as area is increased than islands. In this respect, the isolated mountains behave as true islands. The number of en- demics is low on the islands (never exceeding 5 percent of any flora), however; and the island floras are over- whelmingly dominated by species with no apparent modifications for long-range dispersal. Furthermore, the east- ern mainland has exerted a far greater influence on the flora and the vegetation of the islands than has the western mainland, despite the fact that the former is downwind of the islands. Thus, evidence from endemics, dispersal ecology, and sources of the floras suggests that the isolated mountains have not acquired their full floras by long-range dispersal. We conclude that although the floras of the islands have many insular characteristics, they were less isolated in the relatively recent past than at the present. The island floras do not appear to be in equilibrium in the sense that immigrations equal extinctions. The biogeography of disjunct segments of similar habitat has intrigued biologists since the days of Charles Darwin (1859) and A. R. Wallace (1880). Their pioneering obser- vations were based primarily on oceanic is- lands, but others have analyzed the biology of such habitats as caves (Culver, Holsinger, and Baroody 1973), woodlots (Curtis 1956), fresh water lakes (Barbour and Brown 1974), and isolated patches of herb land in high-elevation forests (Vuilleumier 1970). The appeal of islandlike environments to biologists is partially explained by the fact that complete inventories of selected taxa can be prepared for several disjunct points in a reasonably short time. Furthermore, is- land systems are ideally suited for the anal- ysis of such dynamic processes as dispersal, competition, and evolution. Basic principles of community structure and trophic dynam- ics also appear to have been better demon- strated and more easily studied in island systems than in larger, more heterogeneous environments (Lindeman 1942, Simberloff and Wilson 1970, Brown 1971a, Heatwole and Levins 1972, and MacArthur, Diamond, and Karr 1972). In this paper we consider the vascular plant floras of islandlike enclaves of mesic environment on high mountains in the deserts of the Great Basin. In the strictest sense, these high mountains are less isolated than oceanic islands, since dispersing prop- agules or their carriers may rest in the desert and survive to move on again. Also, species of the mountain islands could evolve (and apparently often have) from the floras of the unfavorable environments that sepa- rate the islands (Billings 1977). Furthermore, evidence suggests that at varying times in the Pleistocene many of the islands were connected by vegetation similar to that now confined to the slopes of the mountains (Wells and Jorgensen 1964 and Wells and Berger 1967). Nevertheless, the tops of the high mountains of the arid West provide disjunct patches of habitat that may have much in common with real islands. 'Department of Botany and Range Science, Brigham Young University, Provo, Utah H4WI2 'Department of Biology, Allegheny College, Meadville, Pennsylvania 16335. 81 82 GREAT BASIN NATURALIST MEMOIRS No. 2 Methods This paper is based entirely on published floras or checklists of workers who have collected extensively on specific mountain ranges. We utilize 9 floras from the more- or-less continuous mountain systems that flank the Great Basin on the west and east and floras or checklists for 15 mountain ranges in or near the Great Basin (Fig. 1, Table 1). We have assumed that the floras of the relatively continuous flanking moun- tain systems (the Cascade-Sierra system in California and the Teton-Wasatch-High Plateau system in Wyoming, Idaho, and Utah) have long had relatively free access to large floras adapted for life at high ele- vations and thus qualify as mainland floras in the parlance of island biogeographers. The mountain islands have been assigned discrete boundaries which are defined by the 7500 foot contour line. The size and elevation of these islands and their distance from the mainlands were taken from topo- graphic maps. Island-to-mainland distances were computed by summing the distances Fig. 1. Location of the floras considered. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 83 is* Is II > ^ o a> ,0. ~H — 05 oo m oo cd m 2 5 in h co tj< i— i cd <-h t~- r~ oo in co CO ■"* ^ o c £ « DC < J? PC E .£ O Q r- c o m o> in oo co -*r t- a> (N o* -T as oo" || Is o i II U £ 1°-* 1 c ■£ m -"= 2 t3 «» S Z PC oo on D S3 84 GREAT BASIN NATURALIST MEMOIRS No. 2 across inter-island barriers of desert (areas below 7500 feet) along the shortest route possible from a particular island to the nearest edge of each mainland. It should be noted that our island areas and distances to mainlands do not always agree with those reported by Brown (1971a), Johnson (1975), and Behle (1977), who have used some of the same islands that we have. Those discrepancies arise from the manner in which the mainland and island borders are defined by the sever- al authors. Brown (1971a), for example, combined the White and Inyo ranges, but the flora used in our work (Lloyd and Mit- chell 1973) covers only the White Moun- tains. Johnson (1975) let the lower edge of forest or woodland serve as the edge of his islands, while we have followed Brown (1971a) and used the 7500 foot contour as the island edge. In Johnson's (1975) work, the Pine Valley Mountains were considered to be part of the mainland, but our criteria dictate that those mountains be considered an island. As noted elsewhere in this symposium (West et al. 1977), distance to the nearest mainland is a weak ecological variable in the Great Basin, since each mountain range has probably received migrant species from both mainlands. We measured the width of valley barrier between each mountain sys- tem and both mainlands in an effort to ob- tain a better understanding of the bio- geographic consequences of distance. Johnson (1975) and Harner and Harper (1976) demonstrated that habitat diversity exerts a strong influence on diversity of birds and vascular plants, respectively. John- son (1975) used plant criteria to quantify habitat diversity for birds on Great Basin mountains, but his criteria for habitat diver- sity would lead to circular logic if they were used to help explain plant diversity. Conceivably, one could devise a habitat di- versity measure based on physical character- istics of the sample areas, but a useful mea- sure would probably require more information about individual mountain ranges than is now available. Accordingly, we have used only area, elevation, and loca- tion in our analysis of factors controlling plant diversity. The component species of each checklist have been individually considered for in- clusion in our study. We have eliminated species from the checklists which are not known to occur above 7500 feet. Species that are potentially able to survive and re- produce in desert environments have also been excluded. This latter criterion was used to improve the likelihood that the is- lands considered are at least currently func- tioning as islands. We experienced difficulty in rigidly applying this last criterion, since some species which occur above 7500 feet along the eastern edge of the Great Basin do not extend above that elevation in the Sierras. We have included all species which occur above 7500 feet on the eastern edge of the Great Basin (that do not tolerate deserts) but normally occur below that ele- vation on the western mainland. For each species included in the study, we have noted lifeform, likely means of dis- persal, and geographic range. The lifeform categories recognized are: 1) annual, 2) per- ennial forb, 3) perennial graminoid, 4) shrub, or 5) tree. Categories of dispersal in- clude: 1) mega wind, 2) miniwind, 3) stick- tight, 4) fleshy fruit, or 5) no apparent mod- ification. Species were placed in the following groups with respect to geographic range: 1) occurring on both mainlands, 2) confined to the western mainlands and a few isolated mountains, 3) confined to the eastern mainlands and a few isolated moun- tains, or 4) known only from one or a few mountains in this study. Because the authors of the several checklists were uneven in their treatment of taxa of subspecific rank, we have ignored such taxa. It will be recognized that many arbitrary decisions are required to classify all of the species in respect to the foregoing charac- teristics. We have followed the lifeform classification given in the index of Hol- mgren and Reveal (1966), except that we 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 85 have separated annual plants from perennial herbs. We consider megawind propagules to be dust-like seeds (as in orchids) and seeds with large plumose appendages (as in milk- weeds) that can be expected to be regularly transported over a mile by wind. Miniwind propagules are considered to include such fruits as winged utricles of some chenopods, samaras of maples, grass caryopses that have large surface-to-volume ratios, and winged seeds of conifers. Normal dispersal distance of miniwind propagules is probably no more than a few yards. The sticktight category includes "hitchhiker" fruits such as those of Xanthium, Arctium, Circaea, and Bidens which are presumably adapted for dispersal on the fur or feathers of vertebrates. Under the heading of fleshy fruits, we include drupes, pomes, berries, and fleshy cones such as those borne by Juniperus. We as- sume that such propagules appeal to and are often dispersed by birds. Propagules des- ignated as having no modifications for dis- persal are produced by a great variety of dry-fruited species in which seeds are rela- tively large, have a small surface-to-volume ratio, and are without wings or plumose ap- pendages. The categorization of individual species according to geographic range also present- ed difficult problems. Once the floras were recorded on computer cards, the species were separated into the four floristic groups previously mentioned. Examination of the lists thus compiled demonstrated that some of the species that supposedly occurred only on western mainlands did in fact also occur infrequently on the eastern mainlands, even though they were not encountered on any of the checklists. In like manner, some spe- cies on the list of taxa found only on check- lists from the eastern mainlands are known to occur (usually sparingly) on the western mainland. Finally, species that occur on is- land checklists but not on mainland lists are rarely local endemics, but are instead north- ern or southern species or uncommon main- land species that have reached some of the isolated mountains. Despite these defi- ciencies of the geographic range lists, we have used them for certain analyses that would have been otherwise impossible to make. We have used Holmgren and Reveal (1966) as our nomenclatural authority for all species occurring in the Great Basin. No- menclature of species that occur in Califor- nia but do not occur in the Great Basin fol- lows Munz and Keck (1959). Species mentioned that occur to the south of the study area but not in California are named according to Kearney and Peebles (1951) or Clokey (1951). Problems of synonymy were largely resolved with the Holmgren and Re- veal (1966) checklist. Results The Study Areas Our floristic samples are drawn from 6 states and from areas ranging in size from 1 to 3,630 square miles. The mainland floras are distributed across a north-south gradient of about 450 miles in the west (3 floras) and 600 miles in the east (6 floras). The 15 is- lands are geographically centered on the Great Basin and are spread across more than 400 miles of distance in both north- south and east-west directions (Fig. 1). Max- imum elevation varies from 14,495 to 9105 ft above sea level among mainland areas and from 14,246 to 8235 ft among islands (Table 1). Unfortunately, few climatological stations are maintained at high elevations in the re- gion. The few data that are available sug- gest that the climates of eastern and west- ern mainlands are somewhat similar in respect to annual precipitation and poten- tial evaporation at comparable elevations, while the island areas tend to receive less precipitation and to experience greater po- tential evaporation than either mainland. Conditions conducive to aridity appear to be maximal on the more southerly of the mountain islands considered (United States Department of Interior 1970). GREAT BASIN NATURALIST MEMOIRS No. 2 The Flora A total of 2,225 different species occur above 7500 ft elevation in the 24 floras considered in this paper. Approximately 27 percent of those species occur on both mainlands and on occasional mountain ranges between the mainlands. Some 29 percent of the species appear on the west- ern but not the eastern mainland, and roughly 30 percent of the species are repre- sented on the eastern but not the western mainland. The remaining species (about 14 percent) were recorded only on island checklists (Table 2). Species representative of those occurring on both mainlands include the following: Aconitum columbianum Nutt. Balsamorhiza sagittate. (Pursh) Nutt. Carex aurea Nutt. Carex lanuginosa Michx. Elymus glaucus Buckl. Epilobium angustifolium L. Equisetum arvense L. Fritillaria atropurpurea Nutt. Geum macrophyllum Willd. Glyceria eUita(Na.sh) A. S. Hitchc. Hackelia floribunda (Lehm.) I. M. Johnst. Lonicera involucrata (Rich.) Bank Qsmorhiza chilensis Hook. & Am. Populus tremuloides Michx. Pinus ponderosa Laws. Purshia tridentata (Pursh) DC. Ribes cereum Dougl. Sitanion hystrix (Nutt.) J. G. Smith Thalictrum fendleri Engelm. Viola adunca J. G. Smith Table 2. General distributional character- istics of the flora considered. Total species Species occurring on checklists from both mainlands Species confined to western mainland or occurring on western mainland and some islands but not on eastern mainland Species confined to eastern mainland or occurring on eastern mainland and some islands but not on western mainland Species recorded only on islands 2.225 613 678 288 Species confined to the western mainland or that occur on the mainland and a few iso- lated mountains include the following: Agropywn pringlei (Scribn. & Sm.) Hitchc. Allium obtusurn Lemmon Artemisia douglasiana Bess. Bromus breviaristatus Buckl. Carex amplifolia Boott Carex tahoensis Smiley Cheilanthes gracillima D.C. Eaton Cryptantha mohavensis (Greene) Greene Hulsea brevifolia Gray Libocedrus decurrens Torr. Mimulus torreyi A. Gray Oryzopsis kingii (Bol.) Beal Pinus jefferyi Grev. & Balf. Populus trichocarpa Torr. & Gray Prunus emarginata (Dougl.) Walp. Sequoiadendron giganteum (Lindl.) Stipa californica Merr & Davy Taxus brevifolia Nutt. Trifolium andersonii A. Gray Tsuga mertensiana (Bong.) Carr. Species confined to the eastern mainland or to that mainland and a few islands are rep- resented by the species listed below. Abies lasiocarpa (Hook.) Nutt. Acer grandidentatum Nutt. Balsamorhiza macrophyllum Nutt. Besseya wyomingensis (A. Nels.) Rydb. Calamagrostis scopulorum M. E. Jones Ceanothus martini M. E. Jones Chlorocrambe hastata (S. Wats.) Rydb. Clematis columbiana (Nutt.) Torr. & Gray Erigeron ursinus D.C. Eaton Geum rossii (R. Br.) Ser. Hierochloe odorata (L.) Beauv. Mertensia arizonica Greene Moldavica parviflora (Nutt.) Britton Orthocarpus tolmiei Hook. & Arn. Pinus edulis Engelm. Picea pungens Engelm. Primula parryi A. Gray Quercus gambelii Nutt. Ribes wolfii Rothrock Thermcypsis montana Nutt. Species occurring on the checklists of some of the isolated mountains but on neither mainland include local endemics as well as more widespread species that penetrate our area from primarily northern or southern floras. Representatives of each of these groups are listed below. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 87 Endemics Listed by Location Ruby Mountain Area Castilleja linoides Gray Eriogonum kingii Torr. & Gray Primula capillaris Holmgren & Holmgren Spring Range Angelica scabrida Clokey & Mathias Antennaria soliceps Blake Castilleja clokeyi Pennell Cirsium clokeyi Blake Opuntia charlestonensis Clokey Penstemon keckii Clokey Potentilla beanii Clokey Silene clokeyi C. L. Hitchc. & Maguire Synthyris ranunculina Pennell Tanacetum compactum Hall Toiyabe Mountains Draba arida C. L. Hitchc. Mertensia toyabensis Macbr. Wheeler Peak Eriogenum Holmgrenii Reveal Species Entering from North Castilleja viscidula A. Gray Cymopterus nivalis S. Wats Erigeron watsoni (A. Gray) Cronq. Selaginella selaginoides (L.) Link Species Entering from South Agastache pallidiflora (Heller) Rydb. Antennaria marginata Greene Aqailegia triternata Payson Arenaria confusa Rydb. Eleocharis montana (H.B.K.) Roem. & Schult. Festuca arizonica Vasey Muhlenbergia wrightii Vasey In respect to lifeform characteristics, the floras of the mainlands and islands (both close to and well removed from the main- lands) do not differ significantly (Table 3). We had anticipated that since perennial forbs show a preference for more mesic sites (Harner and Harper 1973) and the is- land habitats appear to be more xeric than the mainlands, such species might be under- represented on the isolated mountains. The data lend no support to that idea. Woody species and graminoides are also uniformly distributed among the floristic groups re- ported in Table 3. The number of annual species is consid- erably higher on the western as opposed to the eastern mainland (Table 3). In fact, if only herbaceous species are considered, Chi- square analysis demonstrates that the num- ber of annual species on the two mainlands departs significantly from random expecta- tions. Also, significantly fewer annual spe- cies occur in the combined flora of the is- lands than on the western mainland, but the island flora does not differ from that of the eastern mainland in this respect. Chabot and Billings (1972) have noted that annual species are more common in the alpine flora of the Sierras than in other alpine floras of North America. Floristic Diversity Considerations Species-area relationships for the total flora and various lifeform subsamples there- Table 3. Lifeform relationships of the floras considered. The criterion for separation of near and far is- lands was a barrier width of less than or greater than 100 miles. The following four islands constitute the "far islands" category: Deep Creek, Jarbidge, Santa Rosa, and Spring. Expected numbers of species in each category (assuming random distribution of lifeform classes among floras) is enclosed by parentheses. Floristic Group Shrubs Lifeform Class Perennial Herbs Forbs Graminoides Annuals Total W. mainlands 27 111 696 226 181 1,241 (28.1) (114.4) (734.9) (214.7) (148.9) E. mainlands 27 119 754 213 139 1,252 (28.4) (115.4) (741.4) (216.6) (150.2) Near islands 30 108 776 204 147 1,265 (28.7) (116.6) (749.1) (218.9) (151.7) Far islands 18 77 440 136 73 744 (16.9) (68.6) Summation (440.6) Chi-Square = 18.285 (128.7) (89.2) (Not a significant departure from random expectations at 12 degrees of freedom and the 0.95 probability level.) 88 GREAT BASIN NATURALIST MEMOIRS No. 2 of are shown for both mainlands and islands in Figure 2. Three generalizations can be drawn from that figure: 1) there are con- sistently more species per unit area on the mainlands than on the islands, 2) floristic di- versity increases faster on islands than main- lands as area increases, and 3) area usually accounts for more of the variation in spe- cies diversity on islands than on mainlands (i.e., correlation coefficients for species-area relationships are usually larger for islands than for mainlands). Observations 1 and 2 have been duplicated in numerous island biogeographic studies (MacArthur and Wil- son 1967) and are commented on here only to emphasize that the isolated mountains under study do exhibit strong similarities with true islands. The third observation may be partially attributable to the classification of a single flora. We have treated the Bryce Canyon flora as mainland, but Figure 2 demon- strates that its flora and lifeform subsamples consistently fall on the species-area trend line for islands and well below the trend line for mainlands. Correlation coefficients for both mainlands and islands would have been improved had we classified Bryce Canyon as an island. The area lies at the southern extremity of the more-or-less con- tinuous system of highlands extending south from northern Utah and along the western AREA ISO MILES) Fig. 2. Species-area relationships for mainland and island floras. Relationships for the total flora and various lifeform subsets thereof are shown. Mainland data are represented by dots; insular floras are shown with triangles. The individual floras are identified in the diagram for total species combined: numbers cor- respond to specific floras identified in Table 1. Subscript c indicates mainland correlation coefficients or regression equations; subscript i indicates island coefficients and equations. S represents number of species and A represents area. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM edge of the Colorado Plateau. We initially considered the habitat breaks along the highland corridor to be short and in- consequential as migration barriers and thus settled on the mainland classification for the area. In retrospect, it seems likely that the narrowness of the corridor has combined with general climatic differences and unusu- al soils to effectively filter out numerous northern taxa that would otherwise be ex- pected in the area. Slopes (Z-values) for the species-area regression lines of Figure 2 are shown as the exponents of area (A) in the equations associated with the figure. The Z-value of 0.11 for total flora on the mainlands is slightly smaller than values commonly re- ported (e.g., 0.12-0.17 by MacArthur and Wilson 1967). The average Z-value of .19 reported for nested quadrats in pinyon- juniper ecosystems of Utah and New Mexi- co (Harner and Harper 1976) should prob- ably not be compared to the Z-values ob- tained for mainlands in this study, since it seems likely that Z-values for nested quad- rats where the largest sample area is only a few acres will always be larger than values for regional floras from areas ranging in size from a few to several hundred square miles. The Z-value of 0.31 for the total flora of islands (Fig. 2) is well within the range of values (0.20-0.35) reported for a variety of kinds of biota on true islands and close to the theoretically expected value of 0.26-0.27 (MacArthur and Wilson 1967). We call attention in passing to the fact that woody plants have flatter species-area re- gression lines than perennial herbs on both mainlands and islands. The flatness of species-area regression lines for mainlands has been attributed to the fact that small sample areas there carry individuals of many species that are poorly adapted to the sample area but nevertheless occur there because vigorous populations of each such taxon exist in nearby, suitable habitats (MacArthur and Wilson 1967). The steepness of species-area trend lines for is- lands is related to at least two factors: 1) decreasing likelihood of an island being col- onized by dispersing taxa as size decreases and 2) increased likelihood of local extinc- tion of small populations on little islands. Brown (1977) reports Z-values of 0.165 for boreal birds and 0.326 for boreal mam- mals on sites of isolated Great Basin moun- tains. He has previously reported a Z-value of 0.428 for boreal mammals, using a more restricted group of species and a different set of mountains (Brown 1971a). Our Z- value for vascular plants on isolated moun- tains thus lies between those for boreal birds, which seem definitely to be in equi- librium on the moimtains (i.e., neither in- creasing or decreasing in respect to number of species per unit area over long time peri- ods), and small boreal mammals, which are believed to be losing species by local ex- tinction faster than new taxa can colonize. Plants in general appear to behave more like mammals than like birds on the moun- tains considered, and perennial herbs yield Z-values that are especially steep and ap- proach the values reported for mammals. Both area and maximum elevation of the mountain ranges were strongly positively correlated with total vascular species on those ranges in this study (Table 4). There was a weak negative correlation between number of species and distance to the near- est mainland. In multiple correlation analy- sis, only area makes a large contribution to the coefficient of multiple determination (R2). Elevation appears to be so closely cor- related with area (r = 0.66) that it brings little new information into the multiple cor- relation analysis. Distance also enters the multivariate equation; but it, like elevation, contributes only slightly over 0.01 to the Revalue (Table 4). The overwhelming dominance of area in the multiple correlation analysis is, in all probability, an illusion. Wyckoff (1973) and Harner and Harper (1976) have demon- strated that both environmental favorability (annual precipitation and/ or soil texture) and environmental heterogeneity (variation in soil characteristics, elevation, and/or ex- 90 GREAT BASIN NATURALIST MEMOIRS No. 2 posure) exert a strong influence on the number of vascular plant species per unit area. However, since area subsumes all of these variables, it alone consistently ac- counts for a highly significant amount of the variation in floral diversity in almost any suite of samples. Unfortunately, data on environmental favorability and hetero- geneity are not available for the sample of mountains considered here. We have thus resorted to the use of the less definitive but nevertheless useful variables of area, eleva- tion, and distance. We commented earlier on the com- plicating effect of two close mainlands in is- land biogeography studies. In order to bet- ter evaluate the influence of distance between island and mainland on floristic di- versity of the islands, we have measured the width of unfavorable habitat separating every island from each mainland. Then, by using only species that appear to be con- fined to one mainland or to one mainland and a few islands (i.e., species common to both mainlands or unique to islands were excluded), we used simple and multiple cor- relation to analyze the relative influence of area, elevation, and distance from mainland on the number of species from either east- ern or western mainlands on the 15 islands. The results (Table 5) show that distance now becomes the major factor influencing the number of western mainland species on the islands. For eastern mainland species, distance is not significantly correlated with number of species in simple correlation analyses, but it makes a sizeable contribu- tion in the multiple correlation analysis. The dissimilar results for species number- distance relationships for species of western or eastern mainland origin may be related to the fact that the islands considered are on the average more distant from the western mainland (149 miles) than from the eastern (117 miles). In any event, the results in Table 5 seem to suggest that use of a single distance (distance to nearest main- land) as in Table 4 may obscure the impor- tance of distance in studies of island floras. We recognize also that a decrease in spe- cies of any given checklist is to be expected as one moves away from the center of the geographical area sampled for the checklist. Such a decrease with distance would be ex- pected even in large continental areas of relatively uniform climate, topography, geo- logical substratum, and geological history and may have nothing to do with dispersal habits of the species. The decrease may re- flect nothing more than the difficulty expe- rienced by locally evolved taxa as they at- tempt to expand their range through established vegetations. Sources of the Flora In this section we consider the question of source of the floras of the isolated moun- tains. How important a contribution do lo- cal endemics make to the floras of the iso- lated ranges? Are the island floras derived Table 4. Factors influencing the number of vascular plant species on the 15 mountain islands. Distance is measured to the nearest mainland area having an elevation over 7500 ft. Factor Area of island Elevation of highest peak Distance to mainland Contribution to Simple Correlation Coefficient of Coefficient (r) with Multiple Number of Species Determination (R2) .879 .777 .668 .014 -.091 .013 Total.799 R = 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 91 equally from eastern and western mainlands, or is one source more important than the other? In respect to endemics, the data suggest that their contribution to the floras of the isolated mountains is comparatively minor. The number of endemics of moderate-to- high elevations appears to be considerably larger on the Spring Range (the Charleston Mountains which Clokey [1951] studied are part of this range) than on any other range considered here. Yet even on the Spring Range, which Clokey (1951) considered to be about five million years old, endemics account for only about 5 percent of the flora above 7500 ft. Endemics account for less than 2 percent of the White Mountain flora (Lloyd and Mitchell 1973). In contrast, plant endemics on many remote oceanic is- lands account for over 50 percent of the flora (Carlquist 1974). Such data force one to conclude that the mountain ranges con- sidered are far less isolated than remote oceanic islands such as St. Helena, the Ha- waiian Islands, or New Caledonia, where the majority of the flora is endemic. In order to evaluate the relative contribu- tion of western and eastern mainland floras to individual islands, we have separated out species unique to western as opposed to eastern mainlands (see Table 2). The rela- tive contribution of uniquely western or eastern species on individual islands is plotted against distance to the respective mainlands in Figure 3. The data demon- strate that the eastern source area con- sistently contributes many more species to the islands than does the western source area. On only one island (the White Moun- tains) does the western mainland contribute a larger percentage of the total flora than the eastern. As will be shown later (Fig. 4), the preeminence of the eastern source area in island floras can be demonstrated for all dispersal types. To further illustrate the relative contribu- tion of the respective mainlands to the is- land floras, we have compiled a similarity matrix for all possible combinations among the 24 floras (Table 6). Various inter- relationships among floras are summarized in Table 7. At first glance, the low sim- Table 5. Factors influencing the number of vascular plant species on the islands when species occur- ring on both mainlands and on islands only are excluded. Width of barrier (distance) separating an island from each mainland has been determined for all islands. Factor Western Mainland Species Simple Correlation Coefficient (r) with Number of Species Contribution to Coefficient of Multiple Determination (R2) Area of island Elevation of highest peak Distance to W. mainland Factor .502 .644 .646 Eastern Mainland Species Simple Correlation Coefficient (r) with Number of Species Total R = .092 .417 .509 .714 Contribution to Coefficient of Multiple Determination (R2) Area of island Elevation of highest peak Distance to E. mainland .590 .306 .137 Total R = .348 .156 .504 .710 92 GREAT BASIN NATURALIST MEMOIRS No. 2 ilarity values seem to indicate little com- monality among floras, but those values must be evaluated in light of the way in which they are computed: for example, the 37 percent similarity value between the Ruby and Deep Creek Mountains represents 223 species common to those ranges. Read- ers are referred to the similarity equation given in the legend for Table 7 for details of computation. Several relationships reported in Table 7 merit attention: 1) internal similarity of the floras from the western mainland is almost identical to the comparable figure for east- ern mainland floras, 2) the island floras are less similar to each other than are floras from either mainland, 3) island floras are, on the average, more similar to eastern mainland floras than to western mainland floras, and 4) even islands closest to the western mainland have slightly closer floris- tic affinities with the eastern, rather than the western, mainland. The second of the foregoing items indicates, as one might ex- pect, that the flora of individual mountain islands tends to be a more random assem- blage of species than is found in individual floras on either mainland. Items 3 and 4 in- dicate that the island floras have been more influenced by the eastern than the western mainland, despite the fact that they lie "downwind" (in this case, the prevailing westerly winds) from the western mainland. This last fact is visually conspicuous in the field since many of the dominant plants of most of the isolated mountain ranges have eastern affinities. Examples of such domi- nant, or at least abundant, plants include the following: Agropijron spicatum (Pursh) Scribn. & Smith Amelanchier alnifolia (Nutt.) Nutt. Amelanchier atahensis Koehne Artemisia arbuscula Nutt. Artemisia tridentata Nutt. 40,— W 30 m Q LU 20 rs= .84 ^ .56 A / A A A A A A A, A 44 _EASJEBfcL A • ^ \^A • 1 1 • -___• • • 1 10 o 300 50 250 100 150 200 250 3 OO WESTERN 200 150 100 so o EASTERN DISTANCE FROM MAINLAND (MILES) Fig. 3. Percent of the flora contributed by species that appear to have immigrated from the western as opposed to the eastern mainland. The western contribution is shown by dots, the eastern by triangles. The revalue represents the correlation coefficient for the curvilinear correlation between percent of species contributed by the western mainland and distance from that mainland. The rw-value is the correlation coefficient for the curvilinear relationship between contribution of eastern species and distance (length of low elevation barrier between the islands and the eastern mainland). Individual islands can be identified in the figure by referring to island-mainland distances in Table 1. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 93 Bromus anomalus Rupr. Caltha leptosepala DC. Ceanothus martini M. E. Jones Delphinium occidentale (S. Wats.) S. Wats. Geranium fremontii Torr. Holodiscus dumosus (Hook.) Heller Juniperus osteosperma (Torr.) Little Lathyrus pauciflorus Fern. Lewisia rediviva Pursh Oenothera eaespitosa Nutt. Pachistima myrsinites(Puish) Raf. Phlox longifolia Nutt. Primula pamji A. Gray Ranunculus jovis A. Nels. Valeriana occidentalis Heller Since others (McMillan 1948 and Major and Bamberg 1967) have speculated about the relative effectiveness of northern and southern migration lanes from the western outliers of the Rocky Mountains in provid- ing species for interior Great Basin moun- tains, we have investigated that problem us- ing the similarity matrix of Table 6. Below we have summarized the relations of four interior ranges in the Basin (Deep Creek, Ruby, Toiyabe, and Wheeler Peak) with three northern sources (northern Wasatch, Mount Timpanogos, and Red Butte Canyon) and two southern sources (Bryce Canyon National Park and Pine Valley Mountains). Average Percent Similarity with Three Two Northern Southern Mountain Range Sources Sources Deep Creek 31.7 21.0 Rubv 29.7 15.0 Toiyabe 25.0 , 18.5 Wheeler Peak 24.3 21.5 The data demonstrate that although both northern and southern routes have fed spe- cies onto the isolated mountains, the north- ern route seems consistently to have been more effective than the southern. The low similarity of the four interior mountain ranges with the East Tintic Mountains and their higher similarity with mountain ranges such as the Jarbidge to the north suggests that migration from the western outliers of the Rockies has been primarily along the northern rim of the Great Basin and south- ward along the north-south-oriented moun- tain ranges rather than westward across the dry basins that separate the ranges of cen- tral Utah and Nevada. That hypothesis is strengthened by the low similarity shown by the East Tintic Mountains with the three northern sources (average similarity of 18 percent). Certain species seem clearly to have reached the interior mountain islands of the Great Basin via the northern route, while others have apparently reached those islands via the southern route. Species representa- tive of each route are noted below. Northern Route Ceanothus velutinus Dougl. WESTERN SPECIES EASTERN SPECIES 150 200 DISTANCE FROM MAINLAND (MILES) Fig. 4. Regression lines relating percent saturation of eastern or western floristie elements on the 15 islands to distance to mainland. Distance is defined as in Fig. 3. Regression coefficients (r-values) larger than .514 are significant at the 0.05 probability level. The b-values are slopes for the regression lines. 94 GREAT BASIN NATURALIST MEMOIRS No. 2 Kalmia polifolia Wang. Ledum glandulosum Nutt. Finns albicaulis Engelm. Rubus parviflorus Nutt. Southern Route ArctosUiphtjlos patula Creene Nieotiinui attenuata Torr. Pivaphullum ramosissimum (Nutt.) Rydb. Pinus aristata Engelm. Pinus ponderosa Laws. Dispersal Ecology Plant dispersal habits in both mainland and island floras are dominated by types which have no apparent modifications for dispersal and types with weak modifications for dispersal by wind (Table 8). For conven- ience, we refer to the latter category as the "miniwind" modification. Species whose propagnles have no apparent modifications for dispersal account for from 50.4 to 53.7 percent of the species in the floras consid- ered in Table 8. Species having miniwind propagules contribute between 28.8 and 33.5 percent of the species. Together, these two dispersal types account for almost 85 percent of the species considered. On the average, species having propagules modified for long-range dispersal by wind (megawind dispersal type) contribute almost 7.5 per- cent of all species in our floras. Fleshy fruited species contribute slightly fewer spe- cies (average 6.1 percent of all species), and species dispersed by sticktights contribute the few remaining species (about 2.5 per- cent). Our data indicate that dispersal types modified for long-range movement (i.e., fleshy fruit and megawind categories) show no tendency to be overrepresented on the remote islands (Table 8). In contrast, Carlquist (1967) has shown that as many as 58 percent of the plant species that reach Table 6. Similarity among the 24 floras as determined with the Jaccard (1912) similarity index. Values report- ed are percent similarity for all possible pairs of floras. Checklist numbers correspond to those assigned to each area in Table 1. Checklist No 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 1 Albion 2 21 Cassia 3 27 15 Deep Cr. 4 17 11 18 Tintic 5 28 16 32 15 Jarbidge 6 17 11 23 13 19 Kaibab 7 16 12 22 17 18 19 Pine V. 8 34 22 41 18 33 20 21 Raft R. 9 25 14 37 13 40 19 15 29 Ruby 10 26 19 21 17 .30 12 14 27 23 Santa Rosa 11 14 8 23 15 16 25 22 19 17 13 Spring R. 12 20 12 29 17 28 20 20 29 27 25 25 Toiyabe 13 19 16 18 11 23 14 14 21 19 22 11 19 Wamer 14 27 15 39 19 31 21 20 33 36 25 22 29 18 Wheeler 15 16 9 25 12 21 19 18 22 24 17 27 31 18 25 White 16 19 12 20 18 17 27 19 21 15 16 25 17 12 23 15 Bryce 17 14 9 17 7 23 11 13 16 21 17 12 17 23 14 17 9 Lassen 18 25 14 32 18 33 21 21 31 32 18 18 23 18 25 21 19 18 Timpanogos 19 24 14 35 15 34 24 21 32 32 18 19 28 19 25 22 18 22 47 N. Wasatch 20 26 17 28 21 32 19 22 32 25 20 17 24 21 23 16 17 17 44 40 Red Butte 21 15 12 18 8 23 11 12 16 19 15 10 19 25 15 18 9 38 18 21 17 Sagehen Cr. 22 11 6 16 6 20 12 11 14 21 10 13 18 15 14 28 9 29 17 24 15 24 Sequoia 23 21 11 31 11 31 24 17 28 30 15 18 22 17 26 20 19 17 37 39 28 16 20 Uinta 24 17 9 24 9 31 17 14 23 29 14 12 20 16 ,20 19 12 21 32 38 26 19 23 35 Yel 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 95 remote oceanic islands are dispersed inter- nally by birds (mostly fleshy fruits). Prop- agules borne externally on birds by virtue of being held in place by barbs or prickles (sticktights) also account for many (fre- quently over 20 percent) of the in- troductions. He found windborne seeds to be poorly represented (usually less than 10 percent of the flora) on all save the closest of the remote islands. He considered ecolog- ical conditions on the island to be strong determinants of the dispersal types that suc- ceeded. In contrast to Carlquist's findings, Hed- berg (1970) found wind-dispersed plants to represent almost 30 percent of the flora above about 7900 ft on the mountains of east Africa. In Hedberg's study, plants dis- Table 7. Floristic similarity relations among the floras considered. The index of similarity used is that of Jaccard (1912). Jaccard's index is computed as follows: C SI X 100. In the equation, C represents the number of species common to the two floras, A is the number of species in flora A, and B is the number in flora B. Areas Considered No. of Floras Involved No. of Comparisons Averaged Average Percent Similarity Western mainland (internal similarity) Eastern mainland (internal similarity) Mountain islands (internal similarity) W. mainland compared with islands E. mainland compared with islands W. mainland compared with E. mainland Four closest islands to W. mainland compared to W. mainland Four closest islands to W. mainland compared to E. mainland Four closest islands to E. mainland compared to W. mainland Four closest islands to E. mainland compared to E. mainland 3 3 30.3 6 15 30.1 15 105 21.0 18 45 15.0 21 90 21.4 9 18 17.3 7 12 18.1 10 24 18.8 7 12 11.4 10 24 21.0 Table 8. Plant dispersal habits of the floras considered. Expected number of species in each category (assuming random distribution of lifeform classes among floras) is enclosed by parentheses. Island groups are defined as in Table 3. Floristic Mega- Mini- Fleshy Stick- No Group wind wind Fruits tights Modification Total W. mainland 93 .358 88 .36 666 1,241 (92.3) (395.8) (75.3) (30.9) (646.7) E. mainland 104 414 77 26 631 1,252 (93.2) (399.3) (75.9) (31.1) (652.4) Near islands 83 415 65 .30 672 1,265 (94.1) (403.5) (76.7) (31.5) (659.2) Far islands 55 249 43 20 377 744 (55.4) (237.3) (45.1) (18.5) (387.7) Summation Chi-Square = 15.501 (Not a significant departure from random expectations at 12 degrees of freedom and the 0.95 probability level.) 96 GREAT BASIN NATURALIST MEMOIRS No. 2 persed internally by birds accounted for only 1 to 2 percent of the alpine flora of east Africa. The relationship between various dis- persal types and island-to-mainland distance is presented in Figure 4. There, we regress percent saturation of species of various dis- persal habits (i.e., the number of species of a given dispersal habit on each island is ex- pressed as a percentage of the number of species of that dispersal habit that would be expected in an area of comparable size on the appropriate mainland) against distance. As expected, the regression lines all have negative slopes, and there is a slight (but statistically nonsignificant) tendency for dis- persal types that are easily dispersed over long distances (megawind and fleshy fruit types) to have regression lines with gentler slopes than are obtained for species that are less likely to be dispersed far from the par- ent plant. Average slope values for western and eastern mainlands and each dispersal type are shown below. Dispersal Type Average Slope Value Megawind Fleshy Fruits Miniwind Sticktight No Modification .03 .05 .08 .(MS .07 The data in Figure 4 also support our earlier conclusion that the eastern mainland has exerted a greater influence on the mountain islands than has the western main- land. Every dispersal type shows greater saturation for eastern species than for spe- cies from the western mainland. Since the number of species originating from each of the two mainlands is roughly equal (see Table 2), the results in Figure 4 suggest that species from the eastern mainlands have been about four times as effective in reach- ing and surviving on the islands as those from the west. On the average island, west- ern species have a saturation value of 8 per- cent, but the comparable value for species from the eastern mainland is 36 percent. The great disparity between correlation coefficients for saturation-distance analyses for eastern and western species in Figure 4 is noteworthy. In five of the six analyses the r-values are much larger for western spe- cies. It seems possible that those values re- flect a differential in age of the two floristic elements on most of the islands. If the Rocky Mountains are much older than the Sierras, as Billings (1977) reports, it is pos- sible that the eastern floristic element has dispersed essentially to its limit and is now poorly related to distance, while the west- ern element is still actively dispersing. Finally, we call attention to a con- spicuous relationship between range limits of species and plant lifeform. Our data demonstrate that woody plants and per- ennial graminoid species are over- represented in the broad-range category (i.e., occurring on both mainlands) and un- derrepresented in the category of species unique to islands (Table 9). Perennial forbs, on the other hand, display a significant ten- dency toward underrepresentation in the broad-range category and over- representation in the island-only class. An- nual species show no significant trends in this respect. It seems possible that the pat- terns observed reflect evolutionary rather than dispersal processes. In general, woody plants and graminoides appear to be ecolog- ically broad niched and to have the ability to become community dominants. In con- trast, many perennial forb genera seem to be narrow niched and to rarely achieve a dominant place in their community. Discussion Mountains as Islands One might expect an island flora to be distinguished from that of the nearest main- land in a variety of ways. As we began this study, it seemed to us that insular floras should display 1) an overrepresentation of species modified in one way or another for 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 97 long-distance dispersal, 2) fewer species per unit area than observed on the mainland, 3) steeper species-area curves than for main- land floras, 4) uneven stocking of species ecologically preadapted for existence on available islands, and 5) higher rates of en- demism than the mainland. Our results demonstrate that the isolated mountains of the Intermountain West satisfy some of our preconceived notions and thus qualify as islands, but they fail to qualify on other counts. The islands do indeed have fewer species per unit area than adjacent mainlands, and species-area trend lines for islands are steeper than those for main- lands (Fig. 2). Although the amount of en- demism is low on the islands (always less than 5 percent), the amount still appears to be higher than on areas of comparable ele- vation and size on the mainlands. Too, there is uneven stocking of species on the islands. The Pine Forest Mountains of ex- treme northwestern Nevada, for example, are stocked by Pinus albicaulis Engelm., the Santa Rosas by Pinus flexilis James, while the Jarbidge and Ruby Mountains to the east and the Sierras to the west have both. The observed distribution pattern for these and many other species [e.g., Abies concolor (Gord. & Glend.) Lindl. and Picea engel- mannii Parry ex Engelm.] seems explainable only in terms of randomness of colonization and/ or extinction (See Critchfield and Al- lenbaugh 1969 for range details for these and other conifers in the Great Basin.) Our expectations relative to an over- representation of species modified for long- range dispersal on the islands in large part failed. The isolated mountains are over- whelmingly dominated by species with no obvious means for being dispersed great dis- tances. Furthermore, there is no tendency for species with modifications for long-dis- tance dispersal to be overrepresented on even the most distant islands (Table 8). Our data do, however, show a weak tendency for percent saturation of poorly dispersed species (i.e., no-modification, miniwind, and sticktight categories) to decline faster and more reliably (larger r-values) with distance than for megawind and fleshy-fruited spe- cies, which are probably more easily dis- persed (Figure 4). Recent literature references demonstrate that at least some of the species that we have classified as unmodified for dispersal are, in fact, highly adapted for dispersal by vertebrate animals. Although we placed all conifers with unwinged seeds in the unmo- dified-for-dispersal category, a recent paper by Vander Wall and Balda (1977) shows that the Clark's Nutcracker regularly dis- perses the seeds of several pines (P. edulis, P. albicaulis, and P. flexilis) in a sublingual Table 9. Plant lifeform relative to the range limits of the species considered. Expected number of spe- cies appears in parentheses in each category. Range Category Woody Plants Lifeform Class Perennial Herbs Forbs Graminoides Annuals Total Species Occurring on both mainlands 94 (70.0) 333 (368.6) 114 (98.9) 72 (75.5) 613 Occurring on one mainland only 140 (151.1) 795 (796.2) 215 (213.6) 174 (163.0) 1,324 Occurring on islands only 20 (32.9) 210 (173.2) Summation Chi-Square 30 (46.5) = 36.020°° 28 (35.5) 288 lificant departure from random expectations at 6 degrees of freedom and the 0.99 probability level.) 98 GREAT BASIN NATURALIST MEMOIRS No. 2 pouch and caches them in soil suitable for their germination and growth. In addition, the Nutcracker is known to occasionally feed on the winged seeds of Pinus aristata and Pinus ponderosa in northern Arizona. Vander Wall and Balda (1977) have evi- dence for the dispersal of seeds over 13 miles in a single flight by the Nutcracker. In California, the Nutcracker regularly feeds on and caches the seeds of Pinus mon- ophylla Torr. & Frem. and Pinus jefferyi as well as Pinus albicaulis and Pinus flexilis (D. Tomback, personal communication). Johnson (1975) suggests that the Pinon Jay and the Band-tailed Pigeon may also be in- volved in long-distance transport of con- iferous tree seed. J. Pederson (personal com- munication) reports that the Band-tailed Pigeon has been taken several miles from the nearest Quercus gambelii in south- eastern Utah with a crop full of unbroken acorns. Staniforth and Cavers (1977) demon- strate that some seeds of two Polygonum species (P. lapathifolium L. and P. pensyl- vanicum L.) retain viability after passing through the digestive tract of the cottontail rabbit in eastern Canada. The foregoing data lead us to suspect that large seeds from the dry fruits of many species will eventu- ally be shown to be dispersed by vertebrate animals. The foregoing discussion is an acknowl- edgement that we have underestimated the number of plant species that are modified for long-range transport on our islands. Nevertheless, the number of species in the no-modification and miniwind categories is so great on the islands that we are still forced to conclude that the vast majority of the species there did not reach those sites by long-range dispersal. Although the high elevation community types may never have been able to survive on the valley floors at any time during the Pleistocene, as Wells and Berger (1967) argue, many of the com- munity components may have been able to migrate directly across valley floors during that period. Also, as Billings (1977) empha- sizes, climatic cooling would have signifi- cantly narrowed the barriers between is- lands. Our discussion of mountains as islands would not be complete without some com- ment on the question of equilibrium of spe- cies number on the islands. Brown (1977) contends that birds are and small mammals are not in equilibrium on isolated mountains in our study area. Are the plants in equilib- rium? It will be recognized that the equilib- rium argument is based on two assumptions: 1) local extinctions do occur, and 2) new in- troductions occur as often as extinctions on each island. Both assumptions are difficult, if not tactically impossible, to test con- clusively. A definitive test would require that we know of every population of every species on every island, and that we mon- itor each island regularly enough (preferably annually) in order to know when a species became extinct or immigrated and became established there. Obviously, such data are not available for any island in our study. As a consequence, any statement about the status of our islands relative to the equilib- rium question must be based on inferences, not facts. With respect to extinctions, there is con- clusive evidence that Pinus aristata and Pinus flexilis coexisted with Abies concolor and Juniperus osteospenna on Clark Moun- tain in southeastern California about 25,000 years ago (Mehringer and Ferguson 1969). Today neither of these pines occurs there. Similarly, Pinus monophylla and Juniperus osteosperma existed on the Turtle Range 14,000 years ago (Wells and Berger 1967), but do not occur there now. The relatively steep species-area curves for herbs (Fig. 2) may indicate extinctions, but we can offer no evidence in support of that possibility. Concerning new immigrations onto the isolated mountains, there are abundant re- cords of exotic species invading at lower elevations (Young, Evans, and Major 1972). Nevertheless, we know of no documented cases of unaided immigrations onto the mountains of species that cannot survive in at least some microsites on the valley floors. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM There is strong evidence that species modified for long-range dispersal are not overrepresented on the islands relative to the mainlands (Table 8). If extinctions and immigrations had been in equilibrium, even since the close of the Pleistocene, one might have expected long-range dispersal types to be at least somewhat overrepresented on is- lands; but even that tendency is not ob- served (Table 8). As noted above, there is a weak tendency for percent saturation of long-distance dispersal types to decline less rapidly against distance from the mainland than for supposedly less well-dispersed taxa. These two bits of evidence lead us to ten- tatively conclude that the flora of the iso- lated mountains is not in equilibrium, even though some species do appear to be mov- ing about in the area. If the islands are not in equilibrium, the extinction rate must be low for all plant groups and especially so for the woody taxa. We draw this inference from the relative flatness of the species-area curve for most plant groups (Fig. 2) in contrast to mam- mals (Brown 1977). Intuitively, this infer- ence seems valid since herbaceous plants as primary producers should be able to main- tain larger populations than their vertebrate consumers. Woody plants (especially trees) would be expected to maintain smaller pop- ulations than their vertebrate consumers, but would have far greater longevity. Tro- phic position and longevity likely have much to do with the relative extinction rate of vertebrates and plants. Plant groups of differing trophic habit (e.g., vascular sap- rophytes and nongreen parasites such as Co- rallorhiza and Orobanche, respectively, ver- sus photosynthetic forms) and longevity should show different extinction rates. We had not expected to find the eastern mainland (Rocky Mountains) floristic ele- ment to be so much more successful than the western mainland (Sierra) element on the Great Basin mountains. As others have noted in this symposium, the Rocky Moun- tain element also dominates the avian fauna (Behle 1977 and Johnson 1977) and the al- pine flora (Billings 1977) of the isolated mountains. The evidence seems to imply that three basic factors have combined to give the Rocky Mountain element an ad- vantage over that from the Sierra. Those factors are: 1) time, 2) geological parent material, and 3) climate. As Billings (1977) has noted, most of the Great Basin mountains are younger than the Rockies and older than the present Sierra Nevada and Cascade ranges. Thus, species from the east have had longer to colonize the isolated mountains than high-elevation taxa from the Sierra, since that flora must have arisen much later than the first. In ad- dition, propagules of species unique to the western mainlands would have had great difficulty establishing themselves on the mountain islands even after reaching them, since most habitats would have already been occupied by eastern taxa. Plants originating at higher elevations on the western mainland could generally be ex- pected to be adapted to acidic soils, since the Sierra Nevada is primarily composed of acidic, igneous rock (Major and Bamberg 1967). Soils on the isolated Mountains, how- ever, have prevailingly basic to circum- neutral soils. Again, taxa from the eastern mainland would have an advantage in colo- nizing the islands, since the western outliers of the Rockies are prevailing formed from calcareous rocks. In this connection, it is significant that Billings (1950) found as- semblages of Sierra plants in the western Great Basin to be confined to acidic habi- tats on hydrothermally altered rocks. Finally, western plants have evolved in an environment that is less continental (i.e., more moist and thermally less variable) than that associated with the isolated mountains of concern or the western outliers of the Rockies. Johnson (1977) considers the cli- matic variable to be highly influential in confining western bird species to the Sierras. We believe that continentality may similarly increase the difficulty of estab- lishment of western plant species that are dispersed to the mountain islands. As in the 100 GREAT BASIN NATURALIST MEMOIRS No. 2 preceding cases, species from the east would be better preadapted for life on the islands. Niche Expansion Brown (1971b) has shown that the altitu- dinal range of a normally low-elevation chipmunk (Eutamias dorsalis) expands up- ward on Great Basin mountains which lack a high elevation congener (£. umbrinus). In the course of our work on isolated moun- tains in the Region, we have observed sev- eral cases in which plant species also dis- play a niche expansion in the absence of normal competitors. Although quantitative data are lacking, we take this opportunity to put such anecdotal evidence as is avail- able on record. An apparent case of niche expansion is presented by Abies lasiocarpa in the Jar- bidge Mountains. There, in the absence of its common coniferous competitors (e.g., Abies concolor, Picea engelmannii, Picea pungens, and Pseudotsuga menziesii (Mirb.) Franco), Abies lasiocarpa plays a major role in forest vegetation from the sagebrush-grass and streambank communities at low eleva- tions to timberline. We know of no other place where this species succeeds in such a variety of habitats. A double zone of Artemisia tridentata oc- curs on mountainsides of Nevada and west- ern Utah. There the species commonly dominates a wide belt both below and above the juniper-pinyon zone. It appears likely that Artemisia has simply moved into a zone that is elsewhere dominated by larger mesophytes such as Quercus gambelii, Pinus ponderosa, or a rich mixture of moun- tain brush species. In the northern Wasatch Mountains, the range of Acer grandidentatum extends manv miles farther north than that of its common associate in the south, Quercus gambelii. In mixed stands of Acer and Quercus, Acer is normally conspicuous only on slope bases and ravine edges. North of the limits of Quercus, however, Acer dominates both slopes and depressions. The phenomenon can be seen with particular clarity in the southwest corner of Cache Valley, Utah. Although Chamaebatiaria millefolium (Torr.) Maxim, occurs on both of the main- lands recognized in this study, it is rarely a conspicuous component of the vegetation on either. On the remote islands, however, Chamaebatiaria is often common and a con- spicuous part of the vegetation. Finally, West et al. (1977) review evi- dence suggesting that the anomalously high upper elevation of the juniper-pinyon zone on many of the isolated mountains of the Great Basin may be attributable to the low diversity of the high-elevation flora and the paucity of well-adapted competitors. They note also that the niche of both juniper and pinyon appears to be severely compressed on the west flank of the southern and middle Wasatch Range where Quercus gam- belii and Acer grandidentata combine to form a dense woodland. Both juniper and pinyon occur in the flora there, but neither is an important part of the vegetation. Adequacy of Checklists In the inception of this study, we were concerned that the checklists on which our work would be based would be too in- complete to give meaningful results. In ret- rospect, we acknowledge that all of the lists are probably incomplete. Undoubtedly, ad- ditional effort will add a few species to some lists and many to others. Nevertheless, the lists have yielded results that seem rea- sonable and defensible. Furthermore, the sample on hand is already sufficiently large to minimize the possibility that new collec- tions will seriously alter species-area rela- tionships or lifeform and dispersal-type spectra for the floras. Management Implications Species-area curves reveal much that should be useful to natural resource man- agers. The curve for trees on islands in Fig- ure 2, for example, suggests that the Santa Rosa Mountains are drastically understocked with trees. Could trees be successfully in- 1978 INTEHMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 101 troduced there to provide shelter for ani- mals or construction materials for man? Since other islands in the study support so many more tree species than that range, we suspect that introduction of one or a few carefully selected tree species into favorable sites would have a high probability of suc- cess there. Species-area curves also have many useful implications for conservation programs for unusual and rare plant and animal species. Managers will find the basic theory relative to rare species and size of reserves nicely capsulized in the following short, non- technical papers: Terborgh (1974 and 1976), Diamond 1976, Whitcomb et al. (1976), and Simberloff and Abele (1976). Johnson (1975) developed a habitat diver- sity index that accounted for a major por- tion of the observed variation in number of bird species on isolated mountains in the Great Basin. Behle (1977) has verified that the index is a useful indicator of bird diver- sity throughout the Basin. Since that index is based on various plant parameters and the presence or absence of free flowing wa- ter, it has relevance to our discussion here. Many of our small, arid mountain ranges in the Intermountain West have only a few acres of complex forest habitat (a prime variable in Johnson's index) in a single loca- tion and but a few score feet of flowing water. Since the index shows that bird di- versity is highly dependent upon such habi- tat, it would seem prudent for developers interested in preserving the natural biotic diversity of the environment to insure that roads, campgrounds, or buildings not in- fringe upon such habitats. Yet, unfortu- nately, our developments often are centered directly on such microenvironmental rari- ties. By so locating developments, we al- most insure that we will lose some and per- haps many plant and animal species from the entire range. The campground at Blue Lake on the Pine Forest Bange in north- western Nevada is a prime example of such faulty planning. With foresight, the devel- opment could have been placed well away from the lake but still in the open pine groves. Water could have been piped to the campground with minimal disturbance to the natural system around the lake. Instead, the current plan places every visitor in a position to disturb the several unusual plant and animal species that perhaps occur at only that spot on the entire range. ACKOWLEDGMENTS We are indebted to Dr. Noel Holmgren (1972) for his exhaustive pioneer work on the biogeography of the Intermountain West. It has been an invaluable aid as we have planned and pondered this paper. Dr. William E. Evenson prepared most of the computer programs used in our analyses. Our research has been supported in part by a grant from the National Science Founda- tion (GB-39272) and another from the Be- search Division, Brigham Young University. A grant from the U.S. Bureau of Beclama- tion through the Utah State Division of Water Besources has helped defray pub- lication costs for this paper. Literature Cited Allred, K. W. 1975. Timpanogos flora. Unpub- lished master's thesis, Brigham Young Univer- sity, Provo, Utah. Arnow, L. A. 1971. Vascular flora of Red Butte Canyon, Salt Lake County, Utah. Privately printed by the author, Salt Lake City, Utah. Atwood, N. D., and L. C. Higgins. 1976. Checklist of plants of Pine Valley Mountains, Utah. Un- published report. Behle, W. H. 1977. Avian biogeography of the Great Basin and Intermountain Region. Great Basin Naturalist Mem. 2: 55 -80. Billings, W. D. 1950. Vegetation and plant growth as affected by chemically altered rocks in the western Great Basin. Ecology 31: 62-74. 1977. Alpine phytogeography across the Great Basin. Great Basin Naturalist Mem. 2: 105-117. Brown, J. H. 1971a. Mammals on mountaintops: nonequilibrium insular biogeography. Amer. Naturalist 105: 467 -478. 1971b. Mechanisms of competitive exclusion between two species of chipmunks (Eutamias). Ecology 52: 305-311. 102 GREAT BASIN NATURALIST MEMOIRS No. 2 1977. The theory of insular biogeography and the distribution of boreal birds and mam- mals. Great Basin Naturalist Mem. 2: 209-227. Buchanan, H. 1973. Checklists of vascular plants of the Albion, Black Pine, Cassia, and Sublette Mountain ranges of southern Idaho. Unpub- lished report. Buchanan, H., and G. T. Nebeker. 1971. Checklist of higher plants of Bryce Canyon National Park, Utah. Weber State College Printing Depart- ment, Ogden, Utah. Carlquist, S. 1967. The biota of long-distance dis- persal. V. Plant dispersal to Pacific Islands. Bull, of the Torrey Bot. Club 94: 129-162. 1974. Island biology. Columbia University Press, New York. Chabot, B. F., and W. D. Billings. 1972. Origins and ecology of the Sierran alpine flora and veg- etation. Ecol. Monogr. 42: 163-199. Clokey, I. W. 1951. Flora of the Charleston Moun- tains, Clark County, Nevada. University of Cali- fornia Press, Berkeley. Critchfield, W. B., and G. L. Allen- baugh. 1969. The distribution of Pinaceae in and near northern Nevada. Madrono 19: 12-26. Culver, D. C, J. R. Holsinger, and R. Baroody. 1973. Toward a predictive cave biogeography: the Greenbrier Valley as a case study. Evolution 2: 689-695. Curtis, J. T. 1956. The modification of mid-latitude grasslands and forests by man, pp. 721-736. In: W. L. Thomas (ed.), Man's role in changing the face of the earth. University of Chicago Press, Chicago, Illinois. Darwin, C. 1859. On the origin of species by means of natural selection (reprint of the first edition). Watts and Co., London. Despain, D. G. 1975. Flora of Yellowstone National Park. Yellowstone Library and Museum Associ- ation, West Yellowstone, Idaho. Diamond, J. M. 1976. Island biogeography and con- servation: strategy and limitations. Science 193: 1027-1029. Gillett, G. W., J. T. Howell, and H. Leschke. 1961. The flora of Lassen Volcanic National Park, California. Wasmann J. Biology 19(1): 1-185. Graham, E. H. 1937. Botanical studies in the Uinta Basin of Utah and Colorado. Ann. Carnegie Museum, Pittsburgh, Pennsylvania, 26: 1-432. Harner, R. F., and K. T. Harper. 1973. Mineral composition of grassland species of the eastern Great Basin in relation to stand productivitv. Can. J. Bot. 51: 2037-2046. 1976. The role of area, heterogeneity, and fa- vorability in plant species diversity of pinyon- juniper ecosystems. Ecology 57: 1254-1263. Heatwole, H., and R. Levins. 1972. Trophic struc- ture stability and faunal change during recoloni- zation. Ecology 53:531-534. Hedberg, O. 1970. Evolution of the Afroalpine flora. Biotropica 2: 16-23. Holmgren, A. H. 1972. Vascular plants of the Northern Wasatch, 4th ed. Privately printed by the author, Logan, Utah. Holmgren, A. H., and J. L. Reveal. 1966. Checklist of the vascular plants of the Inter- mountain Region. U.S. Forest Service Research Paper INT-32. Holmgren, N. H. 1972. Plant geography of the In- termountain Region, pp. 77-161. In: A. Cronquist, A. H. Holmgren, N. H. Holmgren, and J. L. Reveal. Intermountain flora, vol. I. Hafner Publ. Co., New York. Jaccard, P. 1912. The distribution of the flora in the alpine zone. New Phytol. 11: 37-50. Johnson, N. K. 1975. Controls of numbers of bird species on montane islands in the Great Basin. Evolution 29: 545-567. 1977. Patterns of avian geography and speci- ation in the Intermountain Region. Great Basin Naturalist Mem. 2: 137-159. Kearney, T. H., and R. H. Peebles. 1951. Arizona flora. University of California Press, Berkeley. Lewis, M. E. 1971. Flora and major plant commu- nities of the Ruby-East Humboldt mountains with special emphasis on Lamoille Canyon. Re- port to Humboldt National Forest, Elko, Ne- vada. 1973. Wheeler Peak Area species list. Report to Intermountain Region, U.S. Forest Service, Ogden, Utah. 1975. Plant communities of the Jarbidge Mountain complex, Humboldt National Forest. Report to Intermountain Region, U.S. Forest Service, Ogden, Utah. 1976. Flora of the Santa Rosa Mountains, Humboldt National Forest. Report to Inter- mountain Region, U.S. Forest Service, Ogden, Utah. Lindeman, R. L. 1942. The trophic-dynamic aspect of ecology. Ecology 23: 399-418. Linsdale, M. A., J. T. Howell, and J. M. Linsdale. 1952. Plants of the Toiyabe Moun- tains area, Nevada. Wasmann J. Biol. 10: 129-200. Lloyd, R. M., and R. S. Mitchell. 1973. A flora of the White Mountains, California and Nevada. University of California Press, Berkeley. MacArthur, R. H., J. M. Diamond, and J. Karr. 1972. Density compensation in island faunas. Ecology 53: 330-342. MacArthur, R. H., and E. O. Wilson. 1967. The theory of island biogeography. Princeton Uni- versity Press, Princeton, New Jersey. 1978 INTERMOUNTA1N BIOGEOGRAPHY: A SYMPOSIUM 103 MacDougall, W. B. 1947. Plants of Grand Can- yon National Park. Grand Canyon Natural His- tory Association, Grand Canyon, Arizona. Major, J., and S. A. Bamberg. 1967. Some cordille- ran plants disjunct in the Sierra Nevada of Cali- fornia and their bearing on Pleistocene ecologi- cal conditions, pp. 171-188. In: H. E. Wright and W. H. Osburn (eds.), Arctic and alpine en- vironments. Indiana University Press, In- dianapolis, Indiana. McMillan, C. 1948. A taxonomic and ecological study of the flora of the Deep Creek Mountains of central western Utah. Unpublished master's thesis, University of Utah, Salt Lake City. Mehringer, P. J., Jr., and C. W. Ferguson. 1969. Pluvial occurrence of bristlecone pine (Pinus aristata) in a Mojave Desert mountain range. Interim Besearch Beport No. 14, Dept. of Geochronology, University of Arizona, Tuc- son. Millican, M. T. 1969. Transect flora of the Eagle Peak Begion, Warner Mountains, Modoc Coun- ty California. Unpublished master's thesis, Humboldt State College, Areata, California. Munz, P. A., and D. D. Keck. 1959. A California flora. University of California Press, Berkeley. Nebeker, G. T. 1975. Manual of the flora of the East Tintic Mountains, Utah. Unpublished mas- ter's thesis, Brigham Young University, Provo, Utah. Preece, S. J., Jr. 1950. Floristic and ecological fea- tures of the Baft Biver Mountains of north- western Utah. Unpublished master's thesis, Uni- versity of Utah, Salt Lake City. Bockwell, J. A., and S. K. Stocking. 1969. Checklist of the flora of Sequoia-Kings Canyon National Park. Sequoia National History Associ- ation, Three Bivers, California. Savage, W. 1973. Annotated list of vascular plants of Sagehen Creek drainage basin, Nevada Coun- ty, California. Madrono 22: 115-139. Simberloff, D. S., and L. G. Abele. 1976. Island biogeography and conservation: strategy and limitations. Science 193: 1032. Simberloff, D. S., and E. O. Wilson. 1970. Experimental zoogeography of islands. A two- year record of colonization. Ecology 51: 934-937. Staniforth, B. J., and P. B. Cavers. 1977. The im- portance of cottontail rabbits in the dispersal of Polygonum spp. J. Applied Ecology 14: 261-268. Terborgh, J. 1974. Preservation of natural diversity: the problem of extinction-prone species. Bio- Science 24: 715-722. 1976. Island biogeography and conservation: strategy and limitations. Science 193: 1029-1030. United States Department of Interior, Geological Survey. 1970. The National atlas. U.S. Geo- logical Survey, Washington, D.C. Vander Wall, S. B., and B. P. Balda. 1977. Coadaptations of the Clark's Nutcracker and the Pinon pine for efficient seed harvest and dispersal. Ecol. Monogr. 47: 89-111. Vuillemier, F. 1970. Insular biogeography in con- tintental regions. I. The northern Andes of South America. Amer. Naturalist 104: 373-388. Wallace, A. B. 1880. Island life. MacMillan and Co., London. Wells, P. V., and B. Berger. 1967. Late Pleisto- cene history of coniferous woodland in the Mo- jave Desert. Science 155: 1640-1647. Wells, P. V., and C. D. Jorgensen. 1964. Pleistocene woodrat middens and climatic change in the Mojave Desert: a record of juni- per woodlands. Science 143: 1171-1173. West, N. E., B. J. Tausch, K. H. Bea, and P. T. Tueller. 1977. Phytogeographical variation within juniper-pinyon woodlands of the Great Basin. Great Basin Naturalist Mem. 2: 119-136. Whitcomb, B. F., J. L. Lynch, P. A. Opler, and C. S. Bobbins. 1976. Island biogeography and con- servation: strategy and limitations. Science 193: 1030-1032. Wyckoff, J. W. 1973. The effects of soil texture on species diversity in an arid grassland of the eastern Great Basin. Great Basin Nat. 33: 163-168. Young, J. A., B. A. Evans, and J. Major. 1972. Alien plants in the Great Basin. J. Bange Man- agem. 25: 194-201. ALPINE PHYTOGEOGRAPHY ACROSS THE GREAT RASIN W. D. Billings' Abstract.— Alpine vegetations and floras are compared in two transects across the Intermountain Region. The first extends from the Beartooth Mountains in the central Rocky Mountains to the central Sierra Nevada some 1200 km to the southwest. It includes six mountain ranges. The second transect crosses the Mojave Desert from Olancha Peak in the southern Sierra Nevada to Charleston Peak and thence to San Francisco Peaks in northern Arizona. The largest numbers of arctic-alpine species are in the Beartooth and Ruby mountains, indicating migra- tions of these species along the Rocky Mountain cordillera. The lowest numbers of arctic-alpine species are in the central and western Great Basin and in the Sierra Nevada. Sdrensen's Index of Floristic Similarity was calcu- lated for all possible pairs of the nine alpine areas. There is little correlation of floristic similarity with alpine proximity across the Intermountain Region. Rather, any such correlation seems to be in a north-south direction; this is stronger in the eastern part of the region. Insularity and uniqueness of alpine floras seem to increase to- ward the western part of the basin. This is probably due to evolution of alpine endemics from preadapted low- land taxa. The middle-latitude mountains of North America north of Mexico, for simplicity and convenience, may be grouped into four large systems. With few exceptions, the mountains in these systems trend north to south, a fact of considerable importance in the phytogeography of arctic and alpine plants. The four systems are the Appala- chians in the eastern part of the continent, the Rocky Mountains, the Cascades-Sierra Nevada, and last, but not least, the Great Rasin ranges. The latter three systems domi- nate the western third of the continent. In general, the mountain ranges, in their present forms, are younger the closer they are to the Pacific Coast. The Appalachians constitute a very old mountain complex dat- ing from Permian and Triassic times. Much of the large Rocky Mountain system origi- nated in the Laramide Revolution in late Cretaceous and early Paleocene. The oldest basin ranges also rose during the Laramide Orogeny, but most of these ranges, particu- larly in the west, have been upthrust during a period of time from the Oligocene to the Pleistocene. Additionally, the whole basin floor has been uplifted in the Pleistocene. Orogenic activity continues at present. Even though the Sierran batholith is rather old, the present Sierra Nevada is primarily a product of uplift during the Pliocene and Pleistocene (Axelrod 1962 and pers. comm. 1973; Rateman and Wahrhaftig 1966). Fossil lobed oak leaves at 2850 m in the lower al- pine zone on Elephant's Rack, south of Car- son Pass, lend additional evidence of recent Sierran uplift. The high volcanoes of the Cascades are also of similarly recent age. Alpine Islands in the Great Rasin There are some 200 individual mountain ranges within the Great Rasin. Most of these trend in a general north to south di- rection and are separated by broad desert valleys. In the days when mountain ranges were shown on maps by hachures, Dutton described the pattern as similar to an "army of caterpillars marching to Mexico" (Morri- son 1965). These basin ranges, with eleva- tions which vary from about 1800 m to over 4300 m, are by no means alike either Department of Botany, Duke University, Durham, North Carolina 27706. 105 106 GREAT BASIN NATURALIST MEMOIRS No. 2 geologically or botanically. Holmgren (1972) divided the region, on the bases of floristics and geology, into four main divisions made up of 16 sections. The mountain ranges within any one section have certain charac- teristics in common but also there are some rather remarkable ecological differences be- tween mountains within the same section. Holmgren notes that there is greater varia- tion in floristic composition of the alpine zone across the Great Basin from one peak to another than occurs in any other zone. I agree. For many years, biogeographers working in mountain areas have compared isolated mountain ranges and summits to islands (Wallace 1880, Willis 1922). With increased interest in island biogeography (MacArthur and Wilson 1967), several important papers have appeared which provide quantitative information on the geographical relation- ships among the biota on isolated montane "islands." Notable among these are those of Hedberg (1970) on the Afroalpine floras, F. Vuilleumier (1970) on paramo avifaunas of the northern Andes, B. S. Vuilleumier (1971) and Simpson (1974) on paramo floras, Brown (1971) on mountaintop mammals in the Great Basin, and, recently, Johnson (1975) on bird species on montane islands in the Great Basin. MacArthur (1972) also car- ried his theory over to montane islands in the Appalachians in his study of the distri- butions of thrush species. It is notable that two of these papers (Brown's and Johnson's) are concerned with the islandlike distribu- tion of vertebrates in the Great Basin. Nor have plant geographers ignored the island nature of the Great Basin montane islands (P.V. Wells, pers. comm. July 2, 1968, and, of course, Harper et al. in this symposium). Both Brown and Johnson have viewed the Great Basin desert "sea" and its montane is- lands as being bordered on the east and west by large cordilleras, the Rocky Moun- tains and Sierra Nevada, which they desig- nate as "mainlands" or "continents." The desert sea is open to the south as far as the real sea off Mexico. However, to the north, the desert sea eventually diminishes until it is blocked by the jumbled mountain masses of British Columbia which connect the coastal mountains and the Rockies. Johnson (1975) used the lower edge of forest-woodland as the perimeter of his is- lands. This is a real biological boundary. On the other hand, Brown (1971), Harper et al. (this symposium), and I have defined the lower boundaries of the montane islands rather arbitrarily. Harper et al. and Brown have used an elevation of 7500 ft (2286 m) while I have used 9000 ft (2743 m) as an approximation of the extreme lower eleva- tion of alpine plants (Fig. 1). This latter fig- ure is a somewhat liberal estimate of the lower limits of alpine islands in the Inter- mountain Region, but some alpine sites do exist this low, particularly in cirques. Tim- berline is usually higher than this and is of- ten very ragged at its upper limits. Upper timberline is frequently used as the bound- ary between subalpine and alpine vegeta- tion in North America. However, on most mountains of the earth it is not a particu- larly good boundary, and it is not a good boundary on most American mountains ei- ther, including those of the Great Basin. Timberlines almost always exist at much higher elevations than the lowest patches of alpine vegetation. This is often true around glacial valleys, both those with and those without glaciers at the present time. For ex- ample, timberline around the Athabaska Glacier in the Canadian Rockies is fully 675 m above the terminus of the glacier, where in the morainal gravels there are a number of arctic-alpine plant species but no trees. The reasons for this ubiquitous phenomenon are rather simple: those factors which de- fine the lower limits of alpine species are not necessarily those which limit the up- ward distribution of trees. As Figure 1 indicates, even conservative alpine islands in the Great Basin are much smaller and more isolated from each other than the mountain ranges themselves. But these alpine islands have been both larger and smaller in the past than they are at 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 107 120°W 46"N- ■j:r ftMi BJ V,,^ \i ° c,?c V 18W Fig. I. Map of the Intermountain Region showing, in black, areas above 2750 m in elevation. These represent areas which are wholly or partially "alpine" at the present time. The dashed line at the 1850 m contour is an approximate estimate of what could have been the lower limit of alpine vegetation at full-glacial. Alpine regions used in the northern transect are indicated from left to right by letters: P, Piute Pass (Sierra Nevada); W, Pelli- sier Flats (White Mts.); T, Toiyabe Mts.; R, Ruby Mts.; DC, Deep Creek Mts.; B, Beartooth Mts. Those in the southern transect are: O, Olancha Pk and two nearby peaks in the southern Sierra Nevada; C, Spring Mts. (Char- leston Peak); and SF, San Francisco Peaks. 108 GREAT BASIN NATURALIST MEMOIRS No. 2 present. The areas of such islands during glacial and hypsithermal times have had a great deal to do with their present floristics. For example, Simpson (1974) showed that the larger sizes of Andean paramos as they existed at full glacial are more highly and significantly correlated with numbers of plant species per paramo than are the present sizes of each paramo. Also, the dis- tance between paramos at full glacial is al- most statistically significant in regard to present-day floristic richness— but the effect of present distance between paramos upon floristic composition is not statistically sig- nificant. Present-day alpine climates on the Great Basin mountains are quite cold during the winter. In this respect, basin alpine climates are probably comparable in winter temper- ature to alpine areas of the Rocky Moun- tains and the Sierra, although very few al- pine weather data exist to substantiate this. In summer, the basin mountains are far cooler than the desert below. Daytime sum- mer maximum air temperatures are lowest on crests and ridges and nighttime minima are lowest in cirques and canyons; the same relative conditions exist in the Rocky Moun- tains and Sierra Nevada. These Great Basin alpine areas are con- siderably moister than the desert valleys in both winter and summer. However, except for the Ruby Mountains and nearby ranges, they are drier on an annual basis than al- pine regions of the Rocky Mountains or the Sierra Nevada because they receive less snow. But they do receive enough snow to form persistent drifts in the lee of cliffs and ridges. It is such drifts that shorten the al- pine growing season, keep plants well wa- tered, and allow the growth of alpine plants of certain species. The ranges toward the southwestern part of the basin are relatively drier than the rest because they lie in the most extreme part of the Sierran precipi- tation shadow. There is a trend toward more winter snow and summer rain in a traverse across the basin in an easterly di- rection toward the Rocky Mountains. The summer rains in the east are usually in the form of freshening thundershowers, which, I believe, have much to do with the survival so far south of populations of certain arctic species in the Rocky Mountains and eastern basin ranges. These mountains were much colder and wetter during glacial times up to at least 12,000 years BP. Permanent snow was abundant, and many of the ranges had val- ley glaciers. These glaciated ranges include the White Mountains, Toiyabe, Santa Rosa, Independence, Jarbidge, East Humboldt, and Ruby Mountains. In the latter range, valley glaciers emerged from the mountains onto the now sagebrush-covered plains. Even far to the south, there were glaciers on the Spring Mountains2 and San Francisco Peaks. Even though these glaciers were small as compared to glaciers and icefields on the Sierra and in the Rocky Mountains, they did create cirques which are now the refugia for a number of alpine species. The colder, wetter climate also depressed and telescoped vegetational zones and lowered timberlines an estimated 600 to 1200 m (Baker 1970, Loope 1969, Wells and Berger 1969). Even though timberline in itself is neither the only nor the best indicator of al- pine conditions, such a depression would have greatly increased the size of basin range alpine islands and decreased the dis- tances between them. This situation would have increased the chances of establishment of arctic-alpine species by long-distance dis- persal and, in certain instances, by direct migration over tundralike terrain. Since tim- berline depression has been variously esti- mated and since it did undoubtedly vary from one part of the Basin to another, I have compromised by showing in Figure 1 'In botanical literature, the Spring Mountains are frequently, but incorrectly, called the "Charleston Mountains" due mainly to the title of Clokey's (1951) flora of the central portion of the range dominated by Charleston Peak. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 109 what a 900 m depression would do to the sizes and proximity of alpine islands and continents at full glacial. Not only have the alpine islands been larger in the past, they have also been smaller. During the hypsithermal, about 7500 to 4500 years BP, timberlines of Pinus longaeva in the White Mountains and the Snake Range advanced upward to a level about 100 to 150 m above where they now stand (LaMarche and Mooney 1967). Such a long period of warmth and aridity could have eradicated some alpine species by for- est shading or by eliminating their snowy refugia on the higher peaks. The hard rock geological history of the Basin Ranges is a complex and varied one. I shall not go into it here. Suffice to say that sedimentary rocks, and thus calcareous sub- strata, are more abundant in the eastern ranges while the western rocks tend to be igneous or metamorphic, and yet with some dolomites and other calcareous types. To many alpine plants, the chemical natures of these different kinds of substrates are all- important in marginal habitats. Alpine Vegetations and Floras In any study of vegetation and flora, it is important to distinguish clearly between "origin" and "maintenance" of each. Origin is very difficult to pinpoint, and particularly so with alpine floras because they leave al- most no macrofossils. However, the pres- ence of diploids in widespread polyploid arctic-alpine species may play the role of "cytological fossils." The origin of a par- ticular flora on any mountaintop is the re- sult of the interaction of many factors: past geologic events, past climates, migrations of each species, polyploidy, evolution, and even man's activities. Hypotheses are nu- merous; answers are few. While maintenance of an alpine flora and its vegetation is somewhat easier to under- stand, there is still much work involved. One must measure the characteristics of x>th the physical and biological aspects of environment, preferably along meso- and microtopographic gradients throughout the year (Billings 1973). Also, it is necessary to know the tolerance ranges of plant popu- lations in any particular place. This requires many field measurements and much labora- tory experimentation under controlled con- ditions. We have made only a beginning in understanding maintenance of some alpine plant species. Such studies require the inter- action of many people; they are not one- person jobs. Alpine Vegetations In trying to reach at least a partial un- derstanding of alpine phytogeography and ecology in the Great Basin, it is helpful to start by describing alpine vegetations as they now exist from the Rocky Mountains across the basin ranges to the Sierra Ne- vada. A good approach is to look at these vegetations along a northeast to southwest transect from the Beartooth Mountains on the Montana-Wyoming line to the central Sierra Nevada. Such a transect crosses many of the basin ranges, but of particular inter- est are the Ruby Mountains, the Toiyabe Range, and the White Mountains. These represent the eastern, central, and western basin ranges, respectively. Quantitative analyses of alpine vegetation have been made in the Beartooth Mountains (Johnson and Billings 1962), the northern Ruby Mountains (Loope 1969), the White Mountains (Mooney 1973), and the central Sierra Nevada (Chabot and Billings 1972). I do not know of any quantitative alpine veg- etational data from the Toiyabe or Toquima Ranges; perhaps there are some. The floris- tic information and photographs in Linsdale et al. (1952) are of some help, as are per- sonal qualitative observations which I made in 1949. Space does not allow the presentation of those long tables of vegetational composi- tion which do exist; one simply can refer to the publications listed above. However, the alpine vegetation of the Beartooth Moun- 110 GREAT BASIN NATURALIST MEMOIRS No. 2 tains is typically Rocky Mountain alpine with large expanses of alpine tundra in the true sense: Geum rossii turf in mesic sites, Deschatnpsia caespitosa meadow in moist sites, and Carex scopulorum in wet bogs. The crests and ridges are occupied by rather dense stands of cushion and rosette plants with Silene acaulis, Carex elynoides, and many other species dominating. Early and late snowbeds are abundant and have characteristic species surrounding them. Alpine vegetation in the Ruby Mountains lacks the broad expanses of tundra charac- teristic of the Reartooth. However, south of Lake Peak along the divide there is fairly extensive tundra at elevations from 3140 m to 3300 m. This alpine vegetation is charac- terized by Silene acaulis and Carex pulvi- nata. Similar alpine vegetation exists at the same elevation on the north slope of Wines Peak, with Geum rossii and Silene acaulis being the dominants. Carex scopulorum grows in dense stands around alpine ponds. Some of the best-developed alpine vegeta- tion in the Rubies is in the cirque floors. In Island Lake cirque, the vegetation is domi- nated mainly by Erigeron peregrinus, Salix arctica, Caltha leptosepala, Geum rossii, Sib- baldia procumbens, and Polygonum bistor- toicles. Oxyria digyna is common on rocky, moist sites, particularly around snowbanks. One is struck with the remarkable similarity in species composition and site character- istics to the alpine vegetation of the Rear- tooth some 700 km to the northeast. Essen- tially, the alpine vegetation of the Ruby Mountains is a small, only slightly attenu- ated, isolated example of Rocky Mountain alpine vegetation. In strong contrast, the alpine vegetation of the Toiyabe Range, only 250 km south- west of the Rubies, apparently bears little resemblance to that of the northern Rubies or to that of the Rocky Mountains. It seems to be an open, rocky vegetation with a few scattered alpine grasses such as Trisctum spicatum and various species of Draba and Eriogonum, some of which are endemic. There is not the great variety of alpine veg- etation types which one sees in the Rocky Mountains. More vegetational work is needed in the small and little-known alpine areas of the central Great Rasin; further study may change our ideas of these regions of what might be termed alpine desert. Another 150 km farther southwest is the long and high massif of the White Moun- tains. In contrast to the Toiyabe and To- quima Ranges, this has been rather in- tensively studied environmentally, vege- tationally, and floristically. The alpine vegetation has been well-described by Mooney et al. (1962), Mitchell et al. (1966), and Mooney (1973). Over most of the exten- sive alpine area in the White Mountains, the vegetation is extremely sparse; Mooney (1973) reports a plant cover of only 1.5 per- cent at 4175 m on the side of White Moun- tain Peak. Mitchell et al. (1966) say that vegetational cover on windy, gravelly flats on Pellisier Flats, an area of 21 km2 near the north end of the range, rarely exceeds 15 percent. However, on seepage banks and meltwater runs, vegetational cover ranges from 10 to 95 percent. Pellisier Flats at 4100 to 4430 m has active frost polygons and miniature solifluction steps reminiscent of the Reartooth Plateau. The most striking vegetational feature of the White Moun- tains, however, is the sharp distinction in vegetation and flora between the dolomite barrens and granite or quartzite fell-fields. The dolomite barrens are very desertlike and, although in places they may have veg- etational cover up to 12 percent, in other places there are almost no plants. Phlox cov- illei and Eriogonum gracilipes are character- istic species on the dolomite. A granite fell- field at 3870 m had a vegetational cover of 50 percent and was dominated by Trifolium monoense and Koeleria cristata; the first is essentially endemic to the White Mountains and the second is a cosmopolitan species. The alpine vegetation of the White Moun- tains is decidedly unlike that of the Rubies— but both are Basin ranges. There have been several vegetational studies made on alpine vegetation in the 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 111 Sierra Nevada. These vary considerably be- tween the northern end of the range and the southern, where the peaks are higher and the climate drier. Only 48 km south- west of the White Mountain crest and with- in sight of it, is Piute Pass at 3540 m on the Sierran white granite. Here, Chabot and Billings (1972) found the low alpine vegeta- tion covering less than 30 percent of the rocky ground and characterized by Lupinus breweri, Antennaria rosea, and Carex helleri. Other common Sierran alpine species in similar locations on nearby peaks and ridges are Phlox caespitosa, Penstemon davidsonii, Ivesia pygmaea, and Draba lemmonii. At higher elevations, Oxyria digyna, Polerno- nium eximium, and Hulsea algida are con- stant members of the alpine vegetation usu- ally on scree slopes. The most luxuriant vegetation near here is along snowbank meltwater brooks where Dodecatheon jef- freyi, Lewisia pygmaea, and Sedum rosea occur. On very thin granitic soils which dry out after snowmelt, only Calyptridium um- bellatum, the dwarf Koenigia-like endemic Polygonum minimum, and one or two other species can exist. There is a great difference in alpine vegetation here, not only from that of the Beartooth but also from that of the White Mountains on the near horizon. Alpine Floras While there are a few floras for whole mountain ranges between the Bocky Moun- tains and the Sierra, e.g., Lloyd and Mit- chell (1973) for the White Mountains, the main sources of alpine floras per se appear to be journal papers or theses. Beferring again to Figure 1, I have assembled at least tentative figures on the numbers of alpine species for the same NE-SW transect along which the trend in alpine vegetation has been described. These figures will certainly be changed with additional collecting and publication, particularly of more volumes of Cronquist et al. Intermountain Flora and Howell's proposed Sierran Flora. The sources are the Beartooth Mountains (John- son and Billings 1962), the Deep Creek Mountains (McMillan 1948), the Buby Mountains (Loope 1969), the Toiyabe Bange (Linsdale et al. 1952), the White Mountains (Mitchell et al. 1966), and the central Sierra Nevada (Chabot and Billings 1972). Addi- tionally, I have included floristic data from a series of three isolated alpine areas across the southern desert region. From west to east these are Olancha Peak and its neigh- bors in the Sierra Nevada (Howell 1951), the Spring Mountains (Clokey 1951), and San Francisco Peaks (Little 1941). Many alpine plant species also occur in the Arctic. As a basis for whether or not our species do, I have checked each against Polunin's (1959) Circumpolar Arctic Flora. Species which are not arctic-alpine may be endemic to a continental region of the middle-latitudes or to the mountain range itself or its near neighbors. While we need much more information on both scores, the absolute figures on arctic and endemic spe- cies in each small alpine area can tell us something about migrations into an area and possibly about the evolution of new species after such migrations from afar or from the arid regions below. Nor should we forget that some endemic species may be relicts or that some widespread arctic-alpine species may have become extinct during- xerothermic times, or that some may never have reached certain alpine areas because of lack of adaptation for long-distance dis- persal. The upper part of Table 1 (the Beartooth-Sierra transect) illustrates immedi- ately that there are many alpine species in the central Bocky Mountains that also occur in the Arctic. In the Beartooth Mountains, 91 out of 194, or almost half the alpine flora, also occurs in the Arctic. This is in contrast to the 24 arctic-alpine species in the Deep Creeks and 47 in the Bubies; but these are somewhat smaller ranges. The composition of the alpine floras of both of the latter ranges, however, is very much a Bocky Mountain-type flora. In contrast, west of the Buby Mountains there is a dra- matic drop in not only the total numbers of 112 GREAT BASIN NATURALIST MEMOIRS No. 2 alpine species in the Toiyabe, the White Mountains, and the Sierra Nevada, but also in the number of arctic-alpine taxa present. The arctic-alpine element consists, in these ranges, mostly of such easily dispersed, ubiquitous species as Oxyria digyna, Trise- tum spicatum, Cystopteris fragilis, and An- drosace septentrionalis. Even common spe- cies such as Silene acaulis are missing, not to mention relatively rare taxa such as Koe- nigia islandica, Saxifraga flagellaris, or Phippsia algida. Widespread arctic-alpine species such as Silene acaulis, Polygonum viviparum, and Salix arctica reach their western limits in the Great Basin in the Ruby Mountains and do not reappear in the Sierra Nevada (Loope 1969). Since these species are not that particular in their envi- ronmental requirements (there are apparent places in the Sierra Nevada where they could grow), they either must be poorly adapted to long-distance dispersal or there is or has been an advantage to migrating north or south along the old tried and true Rocky Mountain pathway. Those arctic- alpine species in the Sierra Nevada most likely came down from the north in glacial or cooler times during the Pleistocene. Mi- gration across the Great Basin appears to have been a chancy thing even at full- glacial. Only species with propagules easily dispersed by wind or birds appear to have made it to the central Great Basin moun- tains. Even after possible establishment, some alpine species may have become ex- tinct on the Great Basin peaks in dryer, postglacial times. Axelrod (1976) also sug- gests extinctions of moist environment sub- alpine conifers and alpine species in the Great Basin during these xerothermic epi- sodes. I believe that extended post-glacial dry periods also could have eliminated such dry-site arctic-alpine species as Silene acaulis and Saxifraga cespitosa in the Sierra Nevada where they do not occur today. This suggestion, of course, must assume that these species migrated into the Sierra Ne- vada, probably from the north, during fa- vorable Pleistocene times. Looking at the southern transect in Table 1, the same dearth of arctic species is ap- parent in the Olancha Peak group at the southern end of the Sierra and particularly so on Charleston Peak in the isolated Spring Mountains where only 9 arctic species are known to occur. In contrast, only Table 1. Total numbers of true alpine species and those which also occur in the Arctic (arctic-alpine) in two transects across the Intermountain Region. Data from several sources. Location Total No. of Alpine Species No. of Arctic-Alpine Species Percent of Arctic-Alpine Species Northern Transect Beartooth Mts. Deep Creek Mts. Ruby Mts. Toiyabe Range White Mts. (Pellisier Flats) Piute Pass (Sierra) Southern Transect Olancha Pk., etc. (Sierra) Spring Mts. San Francisco Pks. 194 80 102 91 47 24 30 47 25 11 23 10 21 4 10 13 13 9 23 19 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 113 350 km to the southeast, the San Francisco Peaks have at least 19 arctic species in spite of their isolation. Also, there is a distinct Rocky Mountain cast to their alpine flora. A possible pathway for migration into this Pleistocene volcanic mountain may have been at full-glacial through the Arizona White Mountains to the southeast, or the source may have been through the Wasatch Plateau to the north. Again, we come to the origin versus maintenance problem: these arctic species would not remain in the vol- canic cinders or the glacial cirques of this isolated mountain, no matter how they got there, except for the summer thunder- showers reinforcing the winter snows as a source of water. Much more poorly known are the endem- ics of all of the above mountain ranges. Al- pine endemics are much less common in the central Rocky Mountains than in the Sierra Nevada. This may be due to lack of isola- tion along the Rocky Mountain system. The Sierra Nevada as a whole has many endem- ic alpine species but I do not know how many of these may be narrow endemics re- stricted to the two small regions listed in Table 1. The White Mountains have several alpine endemics, and since they cover so much less area than the Sierra Nevada, the chances of these being narrow endemics are better than in the latter range. The same fact holds true for the isolated ranges of the central Great Basin. Some of these latter endemics may also prove not to be so nar- row, except edaphically, upon further ex- ploration. For example, consider Primula nevadensis (Holmgren 1967). The isolated Spring Mountains, with an alpine flora of only 39 species at most (some of these are probably subalpine), has at least 4 endemic alpine species. It seems that alpine endem- ism tends to be more common in relation to total alpine floras on isolated, nonvolcanic peaks and ranges in the southwestern part of the Intermountain Region; this includes the southern part of the Sierra Nevada. The real answer to the question of distribution of endemics lies in more collecting in alpine environments and systematic description of the taxa. One way of comparing alpine floras is to use Srirensen's (1948) Index of Floristic Sim- ilarity. This is expressed as: IS. X 100 1/2 (A + B) where, ISS = Sdrensen's Index of Similarity c = number of species held in common by two alpine areas A = total number of alpine species in Region A B = total number of alpine species in Region B Taking the available alpine floristic data from each of the nine alpine areas on both transects, I calculated SeSrensen's Indices for each possible comparison. These indices are presented as percentage values in Table 2. Also, in the same table, direct distances in km between the alpine areas are shown. When the Indices of Similarity are plotted against distance for each pair of alpine areas, the result is a great deal of scatter in the points. There is little evidence of in- verse correlation of floristic similarity be- tween these alpine regions with distance. Out of 36 possible combinations, only 4 show indices above 30 percent combined with distances below 200 km. These are the Deep Creek— Ruby Mountains in the east- central Great Basin and Pellisier Flats- Olancha Peak, Pellisier Flats-Piute Pass, and Piute Pass-Olancha Peak, all in the Sierra Nevada- White Mountain complex at the ex- treme southwestern edge of the Great Basin. Two others had indices almost equally as high but at distances between 500 and 700 km; these are the Deep Creek Mountains- San Francisco Mountains and Beartooth- Ruby Mountains combinations. The most distant range, the Beartooth, has higher in- dices of similarity with the Ruby Mountains, Deep Creek Mountains, and San Francisco 114 GREAT BASIN NATURALIST MEMOIRS No. 2 Mountains alpine floras at distances of 650 to 1100 km than any of these latter ranges have with the Spring Mountains at distances less than 500 km. The distant Rocky Moun- tains alpine flora as exemplified by that of the Beartooth has as great or greater an in- fluence on those of most of the Great Basin mountain ranges than does that of the much closer Sierra Nevada. There is even a great- er (two to three times) similarity between the alpine floras of San Francisco Moun- tains and the Deep Creek Mountains at a distance of 531 km than there is between the Deep Creek flora and those of the Sierra Nevada and White Mountains, which are equally distant from the Deep Creek Mountains. The rather low indices of similarity in- dicate at least one fact rather clearly: these Intermountain Region alpine areas are very much like newly isolated islands. This is well illustrated along the southern transect from Olancha Peak to Charleston Peak to San Francisco Peaks. With the two gaps being only 225 and 378 km across, respec- tively, the indices of similarity are only 13 and 14 percent. The Olancha Peak group is clearly Sierran, Charleston Peak is unique, and San Francisco Peaks show strong rela- tionships to the Rocky Mountains and Table 2. Sdrensen's indices of floristic similarity (in percent) between alpine floras for all combinations of al- pine areas in both transects. Numbers in parentheses are distances (in km) between the alpine regions. Also, see map in Fig. 1. Nortl lern Transect Pellisier Piute Bear- Deep Ruby Toiyabe Flats Pass tooth Creeks Mts. Range (White Mts.) (Sierra) Beartooth _ 24% 33% 14% 14% 9% Deep Creeks (676) - 39% 22% 19% 10% Ruby Mts. (692) (153) - 20% 14% 11% Toiyabe Range (917) (322) (257) - 21% 16% Pellisier Flats (White Mts.) (1102) (451) (418) (145) - 34% Piute Pass (Sierra) (1167) (515) (483) (217) (72) - Olancha Pk. Group (Sierra) (1223) (547) (539) (290) (169) (129) Spring Mts. (1110) (434) (475) (325) (298) (306) San Francisco Pks. (1086) (531) (668) (636) (660) (668) South em Transect Olancha Pk. Spring San Francisco Group (Sierra' \ Mts Pks. Beartooth 14% 8% 23% Deep Creeks 15% 13% 31% Ruby Mts. 14% 8% 19% Toiyabe Range 13% 18% 21% Pellisier Flats White Mts.) 36% 16% 19% Piute Pass (Sierra) 32% 10% 4% Olancha Pk. Croup (Sierra) - 13% 13% Spring Mts. (225) - 14% San Francisco Pks. (603) (378) - 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 115 mountain ranges far to the north. These af- finities of the San Francisco Peaks alpine flora have also been noted by Moore (1965). After examining the existing floristic data on alpine species across the Great Basin, it is tempting to delineate some alpine phyto- geographic boundaries. Certainly, there is a sharp boundary just west of the Ruby Mountains separating Rocky Mountain types of vegetation and flora from that of the Great Basin. Where this boundary goes to the south, I am not prepared to say. Cer- tainly, it would appear to be west of the Snake Range and perhaps west of the Grant Range; but this is tentative and probably premature. It obviously lies between San Francisco Peaks and the Spring Mountains. Another alpine phytogeographic boundary lies between the White Mountains and the Sierra Nevada and winds north in sinuous curves until it divides the Carson Range on the west from the Pine Nut Range on the east; but several species pay no attention to such an imaginary line and go their way from the Sierra Nevada into the altered an- desites of the Virginia Range (Billings 1950) and the neve cirques of the Pine Nut and Wassuk Ranges (Billings 1954). Again, if edaphic or moisture conditions are right,some species can get to seemingly in- hospitable mountain ranges and stay there perhaps by infraspecific physiological adap- tations (i.e., ecotypes). Possible Origins of Intermountain Alpine FlorAs Alpine floras originate in all mountains by migration and/or evolution; the moun- tains of the Intermountain Region are no exception. Refugia, both glacial and inter- glacial, are also important. As Figure 1 shows, and as Simpson (1974) has demon- strated for the paramos, successful long- distance dispersal is greatly influenced by the sizes and proximity of alpine islands. Such islands were obviously larger and closer during glacial and cooler times. From distributional evidence, only a relatively few arctic-alpine species appear to be able to migrate far: Oxijria digyna, Saxifraga cespitosa, Trisetum spicatum, and Cys- topteris fragilis, for example. Therefore, it is advantageous for ancestral species to be nearby when mountains arise or new alpine areas come into existence by climatic causes. It also seems to be advantageous to be on a north-south migration route such as the old Rocky Mountains, or even the much younger Cascade-Sierran system. The ability to evolve rapidly by adaptive radiation is also advantageous, as for example, in the genus Espeletia in the Andean paramos. Such plasticity, plus pure luck in being near refugia, either edaphic or micro- climatic, also helps. Interglacial refugia are fully as important as glacial refugia. Such interglacial refugia can be mountaintops such as Mount Washington, New Hamp- shire, altered volcanic rocks as in the Vir- ginia Range of western Nevada (Billings 1950), alpine snowbanks, and the moist, cool floors of impervious cirques (Loope 1969). Loope ascribes the presence of many arctic species in the Ruby Mountains today to its nonporous (moist) cirque floors as compared to the porous, dry cirque floors of the Jarbidge Mountains to the north and those of the Snake and Schell Creek ranges to the south. There are other sources of alpine floras in the Intermountain Region. These are the surrounding desert floras as suggested long ago by Went (1948) and experimentally in- vestigated by Klikoff (1966) and by Chabot and Billings (1972) for the Sierra Nevada. The principal findings reported in the latter paper would apply almost as well to the mountain ranges of the Great Basin, provid- ed there are suitable sites for such migra- tion and evolution, and providing there are species and genera of sufficient genetic plasticity near at hand. The derivation of a new alpine flora from a desert or semidesert flora is aided by the following: a. preadaptation of winter annuals and perennials in regard to physiology and morphology, 116 GREAT BASIN NATURALIST MEMOIRS No. 2 b. the ability to migrate upward during favorable climatic periods into new but not altogether dissimilar environ- ments, c. selection for metabolism at low sum- mer temperatures, d. selection for low temperature starch degradation and sugar translocation at night, e. selection of populations and ecotypes which acclimate metabolic-ally rapidly and ideally, and f. selection for flowering and seed-set in short, dry cool growing seasons. These have been elaborated in detail by Bil- lings (1974). There is every reason to believe that such upward evolution of new "alpine-desert" taxa is taking place in the Great Basin. But the rate may be slower than in the Sierra Nevada due to a smaller preadapted flora and also due to the drier and less snow-pro- tected environments on the peaks. In such a case, there could be a trend toward edaphic endemism because other kinds of habitat di- versity are in relatively short supply. Such upward mobility and long-distance dispersal over great distances in a north-south direc- tion seem to be the principal sources of the alpine floras of these isolated montane is- lands in the central Great Basin. Acknowledgments This work has been aided by National Science Foundation Grant BMS72-02356 A01 (General Ecology) for which I express my appreciation. Many people have helped with data and discussion; I thank particu- larly Lloyd Loope, Brian Chabot, Juliana Mulroy, Gaius B. Shaver, and Shirley Bil- lings. Literature Cited \\mh(»1). 1). I. 1962. Post-Pliocene uplift of the Sierra Nevada, California. Bull. Geol. Soc. Amer. 73: 183-198. 1976. Historv of the coniferous forests, Cali- fornia and Nevada. Univ. Calif. Pub. Bot. 70: 1-62. Baker, R. G. 1970. Pollen sequence from Late Qua- ternary sediments in Yellowstone Park. Science 168: 1449-1450. Bateman, P. C, and C. Wahrhaftig. 1966. Geology of the Sierra Nevada, pp. 105-172. In: E. H. Bailey (ed.), Geology of northern California. Ca- lif. Div. Mines and Geol. Bull. 190: 1-508. Billings, W. D. 1950. Vegetation and plant growth as affected by chemically altered rocks in the western Great Basin. Ecology 31: 62-74. 1954. Nevada Trees, 2d ed. Univ. Nevada Agric. Extens. Serv. Bull. 94: 1-125. 1973. Arctic and alpine vegetations: sim- ilarities, differences, and susceptibility to dis- turbance. BioScience 23: 697-704. 1974. Adaptations and origins of alpine plants. Arctic and Alpine Res. 6: 129-142. Brown, J. H. 1971. Mammals on mountaintops: nonequilibrium insular biogeography. Amer. Naturalist 105: 467-478. Chabot, B. F., and W. D. Billings. 1972. Origins and ecology of the Sierran alpine flora and veg- etation. Ecol. Monogr. 42: 163-199. Clokey, I. W. 1951. Flora of the Charleston Moun- tains, Clark County, Nevada. Univ. Calif. Publ. Bot. 24: 1-274. Hedberg, O. 1970. Evolution of the Afroalpine flora. Biotropica 2: 16-23. Holmgren, N. H. 1967. A new species of primrose from Nevada. Madrono 19: 27-29. 1972. Plant geography of the Intermountain Region, pp. 77-161. In: A. Cronquist, A. H. Holmgren, N. H. Holmgren, and J. L. Reveal (eds.), Intermountain flora. Vol. 1. Hafner Publ. Co., New York. Howell, J. T. 1951. The arctic-alpine flora of three peaks in the Sierra Nevada. Leafl. W. Bot. 6: 141-156. Johnson, N. K. 1975. Controls of number of bird species on montane islands in the Great Basin. Evolution 29: 545-567. Johnson, P. L., and W. D. Billings. 1962. The al- pine vegetation of the Beartooth Plateau in re- lation to crvopedogenic processes and patterns. Ecol. Monogr. 32: 105-135. Klikoff, L. G. 1966. Temperature dependence of the oxidative rates of mitochondria in Dan- thonia intermedia, Penstemon davidsonii, and Si- tanion hystrix. Nature 212: 529-530. LaMahche, V. C, and H. A. Mooney. 1967. Altithermal timberline advance in western United States. Nature 213: 980-982. Linsdale, M. A., J. T. Howell, and J. M. Lins- dale. 1952. Plants of the Toiyabe Mountains area, Nevada. Wasmann J. Biol. 10: 129-200. Little, E. L., Jr. 1941. Alpine flora of San Fran- cisco Mountain, Arizona. Madrono 6: 65-96. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 117 Lloyd, R. M., and R. S. Mitchell. 1973. A flora of the White Mountains, California and Nevada. University of California Press, Berkeley. Loope, L. L. 1969. Subalpine and alpine vegetation of northeastern Nevada. Unpublished doctoral dissertation, Duke University, Durham, North Carolina. MacArthur, R. H. 1972. Geographical ecology: patterns in the distribution of species. Harper and Row, Publications, New York. MacArthur, R. H., and E. O. Wilson. 1967. The theory of island biogeography. Princeton Uni- versity Press, Princeton, New Jersey. McMillan, C. 1948. A taxonomic and ecological study of the flora of the Deep Creek Mountains of central western Utah. Unpublished master's thesis, University of Utah, Salt Lake City. Mitchell, R. S., V. C. LaMarche, Jr., and R. M. Lloyd. 1966. Alpine vegetation and active frost features of Pellisier Flats, White Moun- tains, California. Amer. Midi. Naturalist 75: 516-525. Mooney, H. A. 1973. Plant communities and vege- tation, pp. 7-17. In: R. M. Lloyd and R. S. Mit- chell (eds.), A flora of the White Mountains, California and Nevada. University of California Press, Berkeley. Mooney, H. A., G. St. Andre, and R. D. Wright. 1962. Alpine and subalpine vegetation patterns in the White Mountains of California. Amer. Midi. Naturalist 68: 257-273. Moore, T. C. 1965. Origin and disjunction of the alpine tundra flora on San Francisco Mountain, Arizona, Ecology 46: 860-864. Morrison, R. B. 1965. Quaternary geology of the Great Basin, pp. 265-285. In: H. E. Wright, Jr. and D. G. Frey (eds.), The Quaternary of the United States. Princeton University Press, Princeton, New Jersey. Polunin, N. 1959. Circumpolar Arctic Flora. Cla- rendon Press, Oxford. Simpson, B. B. 1974. Glacial migrations of plants: island biogeographical evidence. Science 185: 698-700. Sorensen, T. 1948. A method of establishing groups of equal amplitude in plant sociology based on similarity of species content. Kongel. Danske Vidensk. Selsk. Skr. (Copenhagen) 5(4): 1-34. Vuilleumier, B. S. 1971. Pleistocene changes in the fauna and flora of South America. Science 173: 771-780. Vuilleumier, F. 1970. Insular biogeography in con- tinental regions. I. The northern Andes of South America. Amer. Naturalist 104: 373-388. Wallace, A. R. 1880. Island life. MacMillan and Co., London. Wells, P. V., and R. Bercer. 1967. Late Pleisto- cene history of coniferous woodland in the Mo- jave Desert. Science 155: 1640-1647. Went, F. W. 1948. Some parallels between desert and alpine flora in California. Madrono 9: 241-249. Willis, J. C. 1922. Age and area: a study in geo- graphical distribution and origin of species. Cambridge University Press, Cambridge. PHYTOGEOGRAPHICAL VARIATION WITHIN JUNIPER-PINYON WOODLANDS OF THE GREAT BASIN Neil E. West1, Robin J. Tausch', Kenneth H. Rea1, and Paul T. Tueller2 Abstract.— About 22 percent of the pigmy conifer woodlands of the United States occur within the Great Ba- sin. Only a very few reports of these woodlands exist in the literature. Available reports are either of general de- scriptive nature or specific analysis of vegetation-environmental relationships on one mountain range. In order to better understand basin-wide synecological patterns, a cooperative study was carried out by personnel at Utah State University and the University of Nevada-Reno between 1972 and 1975. Vegetation, landform, geology, and soils data obtained from 463 systematically placed stands on a randomly chosen set of 66 mountain ranges have been used to derive patterns of latitudinal, longitudinal, and altitudinal variation in the floristic diversity in juni- per-pinyon dominated woodlands across the Great Basin. The latitudinal-longitudinal patterns show greatest environmental and floristic diversity on the higher mountain ranges on the southern end of the Central Plateau portion of the study area where the Great Basin-Mojave Desert transition occurs. This is also where the elevational breadth of the woodland belt is greatest. Juniper- pinyon woodlands are largely lacking from northwestern Nevada. The lowest elevations for the type are found in the Dixie Corridor centered in southwestern Utah. The general elevation of these woodlands is highest in the west-central part of the Great Basin and declines both toward the Sierra Nevada on the west and the Wasatch Front-High Plateaus on the east. Use of the equilibrium theory of island biogeography gave incomplete explanations of the diversity patterns ob- served. Certain conceptual and methodical problems forced by this overly simplistic theory are discussed. The best correlations obtained were between species richness and an index of ecotopic diversity. Correlation of Basin-wide patterns of woodland floristics with surficial geology, landforms, and soils is non- discriminatory. However, broad-scale, phytogeographical variations in these woodlands are closely associated with climatic differences. Although much direct climatic data are lacking, it seems likely that the relative contribu- tions of the transitory frontal systems moving inland from the Pacific, continental cyclones developing over the Great Basin, and convectional storms associated with the moist air from the Gulf of Mexico to induce precipi- tation at different seasons are regionally important in the causation of vegetation distribution and composition. The instability of temperature inversions is a likely determinant of the position of woodlands along the north- ern boundary of the type. The Pacific frontal systems break the inversions most readily and are thought to be the major cause for the lack of this vegetation in northwestern Nevada and on exposed mountain ranges along the northern boundary of the type. Such observations provide leads for relevant ecophysiological research to support or reject these notions. The juniper-pinyon woodlands are a ma- siderations of other Great Basin vegetation jor vegetation type of the Intermountain types and their distributions have been West. West, Rea, and Tausch (1975) have made by Billings (1951), Cronquist et al. indicated that about 325,000 km2 (125,000 (1972), Tueller (1975), and Young, Evans, mi2) are involved. About 7.1 million ha, or and Tueller (1976). They all characterized 22 percent of this total, occur in the Great these woodlands as generally covering low Basin. It is the phytogeography of this por- hills or as forming a belt of vegetation at tion of the type that we wish to consider lower elevations on the higher mountains here. with sagebrush-dominated belts both above Brief descriptions of the Great Basin and below, pinyon-juniper woodlands, along with con- Beeson (1974) mapped the typed and de- •Department of Range Science, Utah State University, Logan, Utah 84322. Present address of Rea: H-8 Group Office, Environmental Studies, Los Alamos Scientific Laboratory, University of California, Los Alamos, New Mexico 87544. 'Renewable Resources Center, University of Nevada, Reno, Nevada 89507. 119 120 GREAT BASIN NATURALIST MEMOIRS No. 2 scribed the general longitudinal and latitu- dinal variations from a large set of vegeta- tional data collected on approximately one- third of the mountain ranges found in the Great Basin. In this paper we wish to exam- ine further aspects of this and additional data, expanding the discussion of longitudi- nal, latitudinal, and elevational floristic di- versity (beta diversity, according to Whitta- ker 1975) and propose possible explanations for some peculiarities of latitudinal limits and variations in elevational extent of the type. Methods Data set collection and prior analysis: The data set available involves quantitative data on landform, geology, soils, and vege- tation collected between 1972 and 1975 for 463 stands systematically placed on a ran- domly chosen set of 66 Great Basin moun- tain ranges (Nabi 1978). We defined woodlands as having at least 25 pinyon and/or juniper trees per hectare (10/acre). At least one tree had to be in our mature size-age-form class. These criteria kept the samples from extending too far into ecotones, yet allowed a good coverage of the main juniper-pinyon woodland type. Stands were sampled at regular elevation- al intervals on all of the four major expo- sures where the juniper-pinyon belt existed on each mountain range. We are therefore able to discuss the Basin-wide distribution of these woodlands from fairly objective bases. The full details of the sampling de- sign, data collection, and prior analysis are given elsewhere (Nabi 1978). Supplementary sampling of the Shoshone Mountains added stands in broad canyon bottoms and on secondary slopes of faces along major ridges. These additions were made to sample for one area a greater range of ecotopic heterogeneity than was provided by the regularly placed upland sites on exposures in cardinal directions. In addition to presentation of broad, longitudinal-latitudinal patterns of woodland distribution via a map derived from field- checked space photography, Beeson (1974) grouped the sampled mountain ranges into three first-order divisions, based on floristic composition. The stress was on the major species found on at least 25 percent of the mountain ranges sampled. These major spe- cies were commonly found in most of the stands on the mountain range where they occurred. This approach yielded one group of ranges with moderate floristic diversity in the Central Plateau region that had consid- erable similarities to the foothills and ranges directly adjacent to the Sierra Nevada on the west and the Wasatch Front-High Plateaus of central Utah on the eastern boundary of the basin. Another grouping with low floristic richness was designated for all lower elevation ranges occurring in the generally lower, drier portions of the Great Basin. The third grouping consisted of ranges with the highest floristic diversity. These all occurred along the southern end of the Central Plateau where the Great Ba- sin-Mojave desert transition occurs. These floristic units were related to pat- terns of surficial geology, landform, soils, and climate. Weak correspondences were found at this scale with all environmental factors except climate. The widescale longi- tudinal-latitudinal differences observed were therefore thought to be largely related to current patterns of precipitation and tem- perature. Current analysis methods: Mueller- Dumbois and Ellenberg (1974), in their re- cent textbook on the Aims and Methods of Vegetation Ecology, emphasize that a com- plete understanding of the distribution of plant communities involves a consideration of flora, accessibility, ecological properties of the plants, habitat, and time. Our further analysis here involves an in-depth exam- ination of the total floristic diversity in our samples of Great Basin pigmy conifer wood- lands in an attempt to relate these floristic diversity patterns to likely paleocological influences and to present environmental 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 121 heterogeneity with emphasis on the present climatic patterns through space and time. We also examined the appropriateness of applications of island biogeographical theo- ry (Simberhoff 1974). These analyses fol- lowed a strategy similar to that used by Brown (1971) and Johnson (1975). Data from the 18 mountain ranges sampled at 200 m elevation intervals were used (Nabi 1978). The variables accounted for were to- tal woodland species per mountain range, total mountain area, total mountain area oc- cupied by woodland (taken from our pre- viously published map), mountain height, width of barrier, and an index of ecotopic diversity. The width of the barrier was calculated similarly to Johnson's (1975) approach ex- cept that we considered distance to the nearest of three "continental" areas: The Sierra Nevada, Wasatch Front-High Plateaus of central Utah, and the northern boundary of the Snake River Plains, all de- fined as in Cronquist et al. (1972). The ecotopic diversity where woodlands occurred on each mountain was indexed by the use of an arbitrary, numerical scale (Table 1) taking into account those variables on which we had direct data. The higher the variety of factor conditions, the greater the presumable ecotopic diversity. Untrans- formed data for the variables assessed are presented in Table 2. Correlation coefficients (r) between the variables were then calculated for both the raw and loga- rithmically transformed data. Additional tabular and graphical organi- zations of the data were employed in order to clarify our broad-scale phytogeographical view of latitudinal, longitudinal, and altitu- dinal variation in these woodlands. A series of range-by-range comparisons were made on the basis of total species richness for each range, average stand species richness, and altitudinal distribution of the woodland. Each type of analysis was mapped and compared with each other and the broad- scale climatic patterns for the region. Results and Discussion Overall species richness: A total of 367 vascular plant species were found within Table 1. Point system for ecotopic diversity score. Range of Range of Values Values Possible Observed Exposure Cardinal slopes having woodlands 1 = north only 2 = north and east, etc. Slope Number of slope classes (10% incre- ments) 0-10% = 1, 0-10% plus 11-20 = 2, etc. Elevation Elevations at which woodlands were sampled (200 m intervals) 1800-2000 m = 1: 1800-2000 m plus 2001-2200 m = 2, etc. Landform Five major landforms encountered (valley, foothill, bajada, terrace, mountain) (defined in Nabi 1978) 1-5 1-4 1-5 2-5 1-5 Geology Soils classified to clan (Nabi 1978) Classified to subgroup according to Soil Survey Staff, USDA, Soil Conservation Service (1951, 1960, 1967) Total possible score 1-19 1-24 H\ 1-8 1-16 (10-40) 122 GREAT BASIN NATURALIST MEMOIRS No. 2 the juniper-pinyon woodlands sampled. The number of species in the woodlands on indi- vidual mountain ranges varied from 164 species on the Shoshone Mountains of west- central Nevada to under 30 species on small isolated ranges of western Utah. Figure 1 shows that most of the mountain ranges had fewer than 50 species in their woodlands. Furthermore, the ranking of species num- bers per mountain range frequencies (Fig. 2) shows an expected log-normal distribution over the sampled ranges (May 1975). Many species were found at only one stand on a given mountain range. If the species fre- quency sequence is plotted against the num- ber of stands in which they occur (Fig. 3) the right hand portion of the curve also fol- lows a log-normal form. The frequency val- ues for the rare species (those occurring in one to four stands) are in excess of the ex- pected value (Fig. 4). These high frequen- cies of rarer species are indicative of an over-saturation of relictual species probably remaining from the vegetation migrations resulting from the more favorable climate of the recent geologic past. This is a situa- tion similar to that found by Brown (1971) for small mammals and Johnson (1975) for permanent resident birds in the same area. Our estimates of such oversaturation are conservative since sampling was restricted to upland areas on all ranges except the Shoshone Mountains, which included some broad drainage bottoms and major ridge faces. Judging from the high species density of the Shoshone Range, inclusion of more pronounced topographic situations such as narrow drainage channels and searches of atypical geological outcrops and unusual landforms on the other ranges, in addition to our predetermined sampling scheme, would have yielded more plant taxa within the woodland belt. If the average number of species per stand within a given mountain range is plotted against the total number of species found in the woodlands of that range (Fig. 5) the envelope including the data points il- lustrates that there is a very general in- crease in stand diversity with increasing flo- ristic diversity of the ranges sampled. The variation, however, is high. Except for the ranges with greatest and least diversity, there are few mountain ranges fitting with- Table 2. Untransformed data used to test appropriateness of equilibrium theory of island biogeography. Pigmy Conifer Total Area Total Area Width of Ecotopic Woodland of of Mountain Smallest Diversity Mountain Species Mountain1 Woodland Height Barrier Score Range Encountered (km2) (km2) (m) (km) (See Table 1) Burbank Hills 19 43 43 2384 .36 10 Confusion 32 145 145 2461 39 1.3 Pine Valley 37 143 143 1989 10 18 Black Pine 38 458 113 285.3 30 16 Toana 38 286 263 2117 31 19 White 40 1691 1305 3091 10 19 Goose Creek 42 1691 1305 2295 28 14 Tushar 46 1412 706 3713 10 18 Pilot 48 321 293 3265 20 24 Excelsior 48 187 187 2532 10 13 East Humboldt 49 592 .54 2456 42 18 Schell Creek .54 152.3 1076 3353 12 22 Mineral 65 298 187 2921 10 20 Monitor 69 1908 1.564 2845 35 22 Toiyabe 76 2069 974 3292 42 24 Needle 93 1213 1083 2699 10 40 Highland 94 286 263 2865 15 27 Shoshone 164 519 4.50 3145 42 32 Mean 58.4 821.4 .564.1 2793.1 24.0 20.5 S.D. 33.2 708.9 512.1 466.0 13.2 7.3 'Area above 1800 m. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 123 in a similar range of limited average stand diversity. The majority of the mountain ranges within any total diversity class have a wide variation of average stand diversity. For example, in the interval of 80 to 95 species on a range, the average stand diver- sity varies from 9 to over 25 species. Latitudinal and longitudinal diversity pat- terns: A currently fashionable way of ex- plaining geographical patterns of diversity would be by application of island bioge- ography theory. The results of the correla- tions of total species richness with total mountain area, total area of woodland on the mountain, width of barrier, mountain Fig. 1. Total number of plant species in the wood- land belts on each mountain range ordered from the richest to the most depauperate. 15-30 30-60 60-120 120" 240 NUMBER OF SPECIES ON A RANGE (frequency octoves) Fig. 3. Relationships of stand frequency to species richness. (366 stands of woodland on 66 mountain ranges.) Fig. 4. Number of species ordered by number of stands in which they occurred (frequency octaves). Line represents log-normal interpolation. Fig. 2. Range frequency of total number of plant species in the woodland belts organized by frequency octaves. Line represents log-normal interpolation. Fig. 5. Relationship of average number of species per stand with total number of species in the wood- lands sampled on each of 66 mountain ranges. 124 GREAT BASIN NATURALIST MEMOIRS No. 2 height, and an ecotopic diversity score for each mountain range are given in Table 3. The linear correlations of woodland floris- tic diversity with total mountain area, woodland area on the mountain, width of barrier, and height of mountain range are all low and not related in a statistically sig- nificant sense. There is, however, a signifi- cant relationship of the index of ecotopic diversity with floristic richness. On a log-log basis the picture changes considerably. The correlation between eco- topic diversity and species density is highly significant with little change in value, with or without data from the Shoshone Moun- tains. Species density is significantly corre- lated with total area and woodland area at the 0.05 level without the Shoshone Moun- tains and barely insignificant with them in- cluded. The relationships between other factors shows scattered significance. Ecotopic diver- sity is significantly correlated with total mountain area and woodland area for both analyses. Mountain height and total moun- tain area, but not woodland area, are signif- Table 3. Correlation coefficients (r) between variables. Lower triangular matrix for log-transformed data, up- per for untransformed data. With Shoshone Mountains Species Density Ecotopic Diversity Mountain Height (m) Total Mountain Area (km2) Woodland Area (km2) Width of Barrier (km2) Species Density .7846° ' .3470 .1182 .1669 .1392 Ecotopic Diversity .8570° ° _ .3446 .2349 .3133 -.1355 Mountain Height .4408 .4227 _ .4676 .3390 -.1743 Total Moun- tain Area .4545 .4862° .5326° _ .9265 00 .0068 Woodland Area .4572 .5049° .4620 .8510"° -.1108 Width of Barrier -.3806 -.1691 -.1688 -.0743 -.2325 Species Density Without Shoshone Mountains Total Ecotopic Mountain Mountain Diversity Height (m) Area (km2) Woodland Width of Area Barrier (km2) (km2) Species Density - .8424°° .3303 .3354 .3479 -.2279 Ecotopic Diversity .8521 °° - .2993 .3035 .3658 -.3125 Mountain Height .4120 .3849 - .4995° .3565 -.2580 Total Moun- tain Area .5612° .5219° .5422° - .9272°° .0459 VV oodland Area .5229° .5198° .4596 .8524°° - -.0978 Width of Barrier -.2763 -.3135 -.2413 -.0790 -.2629 - p £ 0.01 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 125 icantly correlated. Total mountain area and woodland area are significantly correlated; however, all these variables are so inter- related that significant correlations are to be automatically expected. The discrepancy between the actual and perfect correlation is probably due to historical accidents such as fire eliminating some woodland from where it has potential on those portions of mountains where climatic unfavorableness exists, as will be explained later. We interpret this failure of width of bar- rier to account for any of the variation ob- served to be at least partially due to a vari- ety of methodical problems created by some simplistic decisions forced by the current development of the theory. First of all, it is impossible to choose one continent contrib- uting flora to the woodlands. Any one mountain range has floristic affinities with all surrounding areas. The mixture could be better described by the degree of affinities to all possible continents. The Mojave desert also contributes to the understory composition. However, because some of the mountain ranges are imbedded in the Mo- jave Desert, we knew of no way to calcu- late barrier width in such circumstances. The overall height of the mountain range probably yields a poor correlation because these woodlands have both boreal-derived and lowland-derived components. Thus, height of connecting ridges or valleys would have a variable effect on what species could have migrated across the Great Basin and have had a route of escape as environments changed. As Johnson and Raven (1970) have previously pointed out, height has inevitable covariation with total environmental diver- sity. Therefore, it is not surprising that the index of ecotopic diversity, which reflects combined physical variables, gives a much better correlation coefficient. The linear relationship of the index of ecotopic diversity with species richness changes if the Shoshone Mountains, with their expanded sampling scheme, are ex- cluded (R2 = .6156 to R2 = .7096). As can be seen from Figure 6, the increased sam- pling on the Shoshone Mountains did not result in an obvious increase in ecotopic di- versity, but it did give an obvious increase in the number of species encountered. This is possibly due to an increased micro- climatic heterogeneity resulting from the sampling scheme used. Some minor changes also occurred for other relationships when Shoshone Mountain data was excluded (Table 3). The influence of the increased sampling scheme on the results with the Shoshone Mountains has been minor with less effect on the log-transformed data than on the lin- ear data. What this appears to show is that, within a reasonably uniform sampling scheme, the floristic diversity encountered within the woodlands is closely related to physical site diversity. This seems to be as much a function of where the plants can survive as it is a function of how many hab- itats exist on a particular range. The primary contributor to the ecotopic diversity index is the number of different 10 20 30 Ecotopic Diversity Score Fig. 6. Regression between ecotopic diversity index and species density in juniper-pinyon woodlands on 17 and 18 mountain ranges in the Great Basin (with and without the Shoshone Mountains included). 126 GREAT BASIN NATURALIST MEMOIRS No. 2 soil types encountered (Table 2). Analyses thus far completed have shown no relation- ship between any specific soil type and any specific vegetation composition on a Basin- wide basis. The diversity of soil and topo- graphic conditions appears to indicate the potential for diversity of vegetation compo- sition but not necessarily which taxa may be involved. This further complicates any at- tempt at an island biogeographical ap- proach. Attempts to map latitudinal and longitu- dinal differences in the equitability com- ponents of diversity were found to be of little value since juniper-pinyon woodlands are almost uniformly dominated by one or two trees plus one or two shrubs. One stand had only three species altogether. Most stands had only trace amounts of plants oth- er than the most common five or six spe- cies. What equitability gradients that oc- curred are mostly elevational and successional, as will be demonstrated later. The species presence-absence approach is deemed more useful at the scale being fo- cused on here (Hume and Day 1974, Schnell, Risser, and Hilsel 1976). Floristic diversity (Fig. 7) is greatest on the higher ranges across the southern end of the Central Plateau portion of the study area where the Great Basin-Mojave Desert transition occurs. In addition to the rela- tionship to the greater present environmen- tal variation (Table 3), we feel that the high diversity there is partially due to historical probabilities of mixing of species with varying abilities of range expansion and sur- vival in the considerable migration of spe- cies which has apparently taken place in and since the Pleistocene, particularly hyp- sithermal time (Antevs 1948 and 1955; Mor- rison 1965; Birkeland 1969; Spaulding 1974; Martin 1963; Martin and Mehringer 1965; Fritts 1965; Cottam, Tucker, and Drobnick 1959; LaMarche 1974; Phillips and Van De- vender 1974; Van Devender 1974; Wells and Jorgenson 1964; Wells and Berger 1967; Elston 1976). These previous papers and analysis of our data favor the view that the pigmy conifer woodlands were once continuous across the Great Basin and northern Mojave Desert re- gions. Subsequent aridity and consequent plant migrations have fragmented these woodlands into the patterns of distribution seen today. The differences in floristic rich- ness are probably more related to rates of extinction (Stebbins 1974 and 1975) than to long-distance dispersal from source biotas on the several possible continents. The rich- ness of the present woodland flora is thus understandably related to the overall envi- ronmental favorability of the various moun- tain ranges at present and in the recent past. Elevational variability: Billings (1951) and Hollermann (1973a and b) have previously commented on the puzzling variation in the elevational extent of Great Basin juniper- pinyon woodlands. We also noted wide var- iation in the Basin-wide elevational limits of this vegetation type. This variation, how- ever, shows a close relationship to the gen- eral topography of the area, with climatic modification being important in the north- ern portions. Our lowest sample in the juniper-pinyon Number of spec Fig. 7. Isolines of species richness in the woodlands hetween the 66 mountain ranges. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 127 woodland was a stand at 1200 m on the south slope of the Beaver Dam Mountains of extreme southwestern Utah. This area is also the topographic low point of the area studied. The highest lower boundaries are found in the White, Excelsior, and Silver Peak ranges in the southwestern portion of the study area, and the Toquima and Mon- itor ranges of the central plateau, some of the highest parts of the topography of the whole study area. The regional distribution of the lowest woodland boundaries are defi- nitely along the eastern and southern boundaries of the study area (Fig. 8). The highest boundary of the woodland belt is much less definite (Fig. 9), largely due to great differences in mountain heights. The upper extreme of these wood- lands in our data set was a 2800 m eleva- tion stand on the west slope of the White Mountains of California. The geographical distribution of upper boundaries is partially correlated to those of lower boundaries. The difference between the actual eleva- tion of the lower limit and upper limit var- ies greatly from aspect to aspect on a mountain range and from range to range over the Great Basin. An average woodland belt width of about 350 m elevational ex- tent occurred on the mountain ranges we sampled. Further analysis of the data shows extremes of belt width ranging from over 700 m in elevation in the White Mountains of California, Highland Peak of Nevada, and the Needle Range of Utah to essentially zero at the northern boundaries of the type. The band of woodland is also extremely narrow in western Utah. These elevational extremes and band widths are graphically portrayed in Figures 10 through 13, which depict transects across the Great Basin. These transects could be combined to make a response envelope. The breadth of these woodlands is generally greatest in the areas where the extremes of elevational extent occur. (Exceptions to this will be explained later). This greater breadth also approximately coincides with and contributes to the higher floristic diver- sity in the Great Basin-Mojave transition discussed earlier. Although undoubtedly some of the eleva- tional variation observed is due to recent historical causes, especially fire and wood ■ 2200 mele'S ED 2100 H 2000 □ 1800 ffl 1600 E3 1200 (Hi m Fig. 8. Isolines of lowest woodland elevation sam- pled. Interpolation between 66 mountain ranges. Fig. 9. Isolines of highest woodland elevation sam- pled. 128 GREAT BASIN NATURALIST MEMOIRS No. 2 cutting (Lanner 1976), there are consistent patterns of elevational extent related to to- pography and climate. The west and east relationships show this most clearly. Wood- lands in the eastern Great Basin are gener- ally lower than those to the west. The woodland belt also narrows considerably at the Wasatch Front. The average elevation and then decreases, while the increases, Fig. 10. Map of the mountain ranges sampled in this study. Lines connect the mountain ranges depict- ed in the following figures. West to east-central transect 4200- 4000- 3800 3600 I i 3400. % 3200- A 3000- 2800 A 3 2600 | 2400- - 2200 : i t 2000 1 1800 . I 1600 " 1400 1200 / 1000 800 600 Fig. 11. Schematic cross section of the altitudinal and latitudinal variations in the woodland belt and its tree species dominance for a west-east transect (see Figure 10) of some of the southcentral ranges sampled. Lines across mountain indicate elevations of adjacent vallev floors. width of the woodland type generally de- creases from south to north. The relative composition of the dominant trees and associated understory varies con- siderably within the belt (Fig. 11-13). The usual case is for pinyon to increase with al- titude and juniper to dominate the lower half of the belt. On many of the higher mountain ranges in the Great Basin, the up- fjA :« W*^ 40C T 38C West to east— southern transect a 90% P«yo« 4 0 00 ? 4 WJW 3800 Sp 3 600- > 3400 S * 320oJ *" 6 f\ 3000 2800- * 2600 1 % 2400 - 2200 A 0 2000 A :. a - a 2 1800 » a '■ * I 1600 * 1200 \ 1000 s. BOO ■* . 1 1 1 1 1 Fig. 12. Schematic cross section of the altitudinal and latitudinal variations in the woodland belt and its tree species dominants for a west-east transect (see Figure 10) of some of the southernmost ranges sam- pled. Lines across mountains indicate elevations of ad- jacent valley floors. North to south transect Fig. 13. Schematic cross section of the altitudinal and longitudinal variations in the woodland belt and its tree species dominants for a north-south transect (see Figure 10) in the Great Basin. Lines across moun- tains indicate elevations of adjacent valley floors. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 129 per part of the woodland belt is a pure pi- nyon overstory. Juniper is more widespread, dominating areas in the north and at lower elevations that do not have pinyon. The quantity of juniper appears to increase with increasing soil moisture stress over most of the Great Basin. The greater proportions of pinyon don't always appear to be purely due to increas- ing precipitation, usual at higher altitudes. Single leaf pinyon appears to have a wider range of moisture tolerance than does juni- per at the higher moisture levels. The in- cidence of pinyon pine with double needles and amounts of juniper increases as the pro- portion of summer rainfall increases from west to east across the Great Basin. The summer dry, extreme western boundary of the basin is the only place where some mountain ranges have single-leaf pinyon but lack juniper. On the west side of the White Mountains there are belts of pure pinyon both above and below juniper (St. Andre, Mooney, and Wright 1965). These two cir- cumstances suggest that although summer soil moisture stress may largely explain the lower elevational limits over most of the basin, the circumstances in the southwestern portion of the study area indicate that there may be temperature regimes unsatisfactory for juniper-pinyon mixes there. St. Andre, et al. (1965) have speculated that lack of competition from extensive for- est vegetation above the pinyon-juniper belt may explain the extreme width and high upper limit of the type on the White Mountains. The general absence of pinyon and juniper on the east flank of the Sierra might be partially due to Ponderosa pine and montane chaparral providing more competition than the sagebrush-grass types typical of the Great Basin Ranges. Similarly, oakbrush-dominated vegetation in the Wasatch Mountains and those near the east- ern boundary of the Great Basin commonly occur above or replace juniper-pinyon woodland vegetation. Could oakbrush pro- vide competition limiting the extent of the pigmy conifer woodlands there? Could the likely upward migrations of juniper-pinyon woodlands and the other conifer belts (La- Marche and Mooney 1972) during the hyp- sithermal have resulted in the elimination of conifer forests at higher elevations of the smaller mountain ranges of the Great Basin? When temperature cooled and precipitation increased in more recent times (LaMarche 1974), pinyon and juniper could have ex- panded with less competition from other tree forms. Are current precipitation- temperature relationships suitable for more competitive species? Unfortunately, little di- rect data exists to answer these questions. Ponderosa pine exists in a few scattered locations on the ranges of southwestern Utah and southeastern Nevada without dominance occurring. In those areas ponde- rosa pine seems to be limited to the most favorable sites, e.g., moist sites along streams or northeastern slopes. Pinyon and juniper compete largely with shrubs at both their upper and lower boundaries within the Great Basin. Until more definitive evidence is in from palynological, dendroclimatological, and woodrat midden studies, we can only go on inferences from current vegetation. Al- though competitive and historical factors are part of the environmental complex, we feel that we can largely explain the current distribution of juniper-pinyon woodlands on the basis of recent and current climatic pat- terns. The effect of mountain height and mass on climatic patterns (the "Merriam Effect," Lowe 1964) has long been recognized by plant geographers. Since its manifestations are greatest in semiarid regions, we related our data on width of woodland bands against the variables of maximum mountain height and area of the mountain mass. Nar- rower belts are generally found on the mountains of smaller area with the narrow- est of the belts generally, but not always, found on those ranges with less than 700 km2 area (Fig. 14). Mountain area above this size has little relation to the width of 130 GREAT BASIN NATURALIST MEMOIRS No. 2 belts. The widest belts are found on the highest peaks and are relatively indepen- dent of mountain area (Fig. 15). A high peak will not, however, necessarily have a wide belt of woodland (e.g., Wheeler Peak, Ruby Mountains, etc.). In circumstances where the belt is wide, the height of the peak dominates, but the location of the peak in the Great Basin is of much more importance. Figure 16 makes a separation of the Great Basin on the basis of where woodland belt widths are either greater or less than 400 meters. There are many peaks in the area of narrow belt width that are as high or higher than those found in the area of the wide belt width, but they are all in the region of greater orographic impact from Pacific Frontal storm systems (Hough- ton 1969). Because of their elevation, the very high ranges of east central Nevada have a climate similar to the ranges more north and west. These climatic controls will now be expanded on. The lower woodland boundaries in the northern half of the study area appear to be related to valley bottom topography (Figs. 10-13). Wernstedt (1960) shows that little difference occurs in summer temperatures between the southern and northern Great Basin, but greatly lowered winter values are encountered as latitude increases. This trend applies best to valley bottom stations. Knowledge of temperature within the woodlands is minimal because few mete- orological stations with long-term records exist. Most of the U.S. Weather Service me- teorological data is from desert valley bot- tom stations where most of the scarce hu- man habitation occurs. Billings (1954) has correlated the occurrence of juniper-pinyon woodlands in northwestern Nevada to ther- mal belts. This led us to suspect that the breadth of the woodland belt all across the northern portions of the Great Basin is at least partially related to the strength of de- velopment and persistence of thermal belts. Analysis of what temperature data that do exist (Fig. 17) shows a strong relationship of fewer degree-days below freezing indexes 2500 3000 MOUNTAIN HEIGHT (metert) Fig. 14. Average woodland belt width (in nearest 100 m increments) in relation to maximum mountain height. 600 900 MOUNTAIN AREA Fig. 15. Average woodland belt width (in nearest 100 m increments) in relation to mountain area above the lower boundary of woodland. Average woodland Fig. 16. Division of the Great Basin on the basis of mountain ranges having an average woodland belt width of greater or less than 400 meters vertical ex- tent. 1978 INTERMOUNTA1N BIOGEOGRAPHY: A SYMPOSIUM 131 for meteorological stations within the wood- land belt, except in the southeastern Great Basin (Pioche and Ursine, Nevada). Degree days above freezing are much higher in most woodland areas than those in the val- ley bottoms where various cold-winter, sem- idesert shrub-dominated vegetation types occur. Juniper-pinyon woodlands are nearly ab- sent north of Interstate Highway U.S. 80 across northwestern Nevada. Critchfield and Allenbaugh (1969) noted the correspondence of the northwestern limits of Pinus mon- ophyUu with the southern boundary of plu- vial Lake Lahontan and speculate as to the lake being a possible barrier to plant migra- tion. The occurrence of pinyon pine on the Virginia and Stillwater Ranges and the fre- quent invasion of juniper in areas that were once under the pluvial lakes of the Great Basin makes this seem unlikely. The corre- spondence of lakes to the pools of cold air and/or lower precipitation now in these ba- sins seems more of an etiological factor in present distribution patterns. Arlo Richardson (personal communi- cation) has mapped the progression of late winter and early spring chill-unit (Richard- son, Seeley, and Walker 1974) accumula- tions across the Great Basin. The valleys of northwestern Nevada have the most rapid progression and resultant ending of winter plant dormancy of any area in the Great Basin. Warming periods are, however, more dramatically punctuated by cold periods brought by Pacific Frontal storms there than elsewhere. This knowledge, combined with existing knowledge of air circulation and storm patterns summarized by Hough- ton (1969), leads us to speculate that a pri- mary reason for the narrowing and absence of the woodland belt in the north, in addi- tion to the latitudinal decrease in solar in- put, is due to the increased frequency of Pacific frontal winds breaking thermal in- versions. This pattern is thought to encour- age earlier plant growth but greater sub- sequent susceptibility of plants to direct frost damage or frost drought in a manner similar to that found for other conifers (Hocker 1956, Newnham 1968). Supporting evidence of the net effect on woodland trees can be seen on space photography where the most northerly and westerly of the mountain ranges in the Great Basin are devoid of pinyon and juniper. In the north central portion of Nevada, ranges such as the Santa Rosa's are almost devoid of con- ifers altogether (Critchfield and Allenbaugh 1969). Ranges to the south and east of the ranges devoid of the woodland contain one or both tree species, implying a lee pro- tection effect, e.g., Spruce Mountain, the southern Ruby Mountains, and the southern most tip of the East Humboldt Range (Fig. 18). In the southern Great Basin the westerly fronts have lost most of their energy; thus their disruptive effect on the development and maintenance of thermal belts is not as frequent. Furthermore, most of the precipi- tation in the southern Great Basin comes from the moist Gulf air masses with rainfall triggered by the lows aloft. As a con- sequence, there are weaker adiabatic lapse rates (decrease in temperature with eleva- tion) and a corresponding smaller increase in precipitation with elevation because of reduced orographic effects. For example, on the west slope of the White Mountains, the Below 0° C Fig. 17. Accumulated degree days cold stress below 0°C for selected woodland (PJ) and valley floor (V) climatic stations across the Great Basin (west on left, east on right). Data obtained from U.S. Weather Ser- vice climatic summaries for years indicated. 132 GREAT BASIN NATURALIST MEMOIRS No. 2 change in precipitation with elevation over the entire woodland belt is only 10 cm (4 inches) (St. Andre et al. 1965). Although west slopes are usually the most mesic and diverse in the Basin, the wood- lands on some of the mountain ranges in our study areas have higher floristic richness on the eastern and southern slopes (Table 4). This apparent reversal of the usual trend is believed to be due to the effects of nearby ranges creating rain shadows and other orographic effects. For example, in the Silver Peak Range most of the moisture must come from lows aloft positioned to the north and east. Johnson (1956) noted that the woodland belt on the south and eastern sides of the nearby Kawich Mountains had the most breadth, highest cover, and largest trees, providing corroborative evidence of this situation. In the southern two-thirds of the study area, the lower boundary of woodlands ap- pear to be more controlled by precipitation than temperature. Degree day values for Pioche and Ursine, Nevada (Fig. 17), are an example of how heat sums do not differ be- tween foothill, woodland, and valley stations in the southeast. In fact, the situation may be slightly reversed over that of the more northerly and westerly pairs of stations. Further evidence of the interaction of temperature and moisture may be observed in the variation of the width of the wood- land belt as related to mountain symmetry. Generally, the north-south axes of Great Ba- sin mountain ranges are longer and the highest peaks are also usually in the center (Lustig 1969). When the range is protected to the west (e.g., Table Mountain in the Monitor Range) (Fig. 19), the widest belts are in the wider, higher, central east-west axis; and the belt-width decreases and aver- age elevation increases toward the north and south ends. This is an apparent re- sponse to the Merriam effect of the greater mountain mass intercepting more moisture at a similar elevation. Extra high mountains such as Arc Dome in the Toiyabes (Fig. 19) appear to effect excessively cold, wet condi- tions for woodland development and the high, central axes have narrower, lower belts primarily on western slopes under such circumstances. We have ignored here the short-term suc- Table 4. Number of species (and number of stands) on each aspect of selected ranges in a west to east belt. Average Mountain Range N E S W Total per stand Virginia 35(2) 17(1) 27(2) -H 50(5) 19 « Pine Nut 32(2) 28(2) 23(2) 32(2) 57(8) 18 Clan Alpine S Desatoya 21(2) 15(1) -H -H 28(3) 12 19(1) 20(1) 15(2) 28(2) 47(6) 15 Needle 31(2) 16(2) 17(2) 24(2) 49(8) 14 30(2) 21(2) 13(2) 26(3) 50(9) 12 55 (13) 53 (16) 33 (15) 43 (12) 92 (56) 9 | WahWah 20(2) 19(2) 17(2) 16(2) 45(8) 10 Mineral 23(4) 18(4) 31(5) 32(4) 62 (17) 9 Stansbury 13(1) 23(2) 12(1) -H 37(4) 12 Tushar 32(2) 17(2) 18(2) 9(1) 48(7) 10 Average 26.3 22.9 18.7 23.3 52.0 10.1 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 133 cessional changes that are known to be op- erating in these woodlands. If we plot by aspect, relative tree cover, richness, and rel- ative equitability against elevation for an example mountain range (Fig. 20), we see that richness is minimal in the closed wood- lands in the middle of the belt where un- derstory has been largely excluded by high relative tree cover. Equitability patterns are more complicated. The approach to equitability used mea- sures how evenly the existing total cover is divided among the species present. For each different total cover and species number a different maximum equitability is possible. To make an equitability comparison be- tween areas of different cover and density, it is necessary to convert the figure to a rel- ative one; in this case the percent of the maximum possible for each site. During suc- cession (West et al. 1975) invasion by juni- per and pinyon shifts in relative cover to- ward more trees. However, before a significant drop in total species number oc- Fig. 18. Effects of exposure to westerly winds and storm tracks on occurrence of the woodland belt at the northern boundary of the type in northeastern Ne- vada. Stippled area is juniper-pinyon woodland, outer (lower) line is 2134 m (7000 ft.) contour, inner (higher) line is 2743 m (9000 ft.) contour. Dots are major mountain peaks: SS = Sulphur Springs Range, EH = East Humboldt Range, Sp = Spruce Mountain, R = Ruby Mountains. (Vegetation map derived from LANDSAT-1 photography. Elevational contours from U.S. Geological Survey maps.) curs, equitability drops. As tree suppression substantially reduces the number of under- story species, equitability rises. Equitability is thus highest at either end of the sere and lowest in the intermediate serai stages. This is contrary to the linear model for xeric- forests proposed by Auclair and Goff (1971). Fig. 19. Interactions of exposure to westerly storm tracks, mountain height, and mountain mass on the variation in woodland belt width and elevation in west-central Nevada. Stippled area is juniper-pinyon woodland. Outer (lower) line is 2134 m (7000 ft.) con- tour. Inner (higher) line is 2743 m (9000 ft.) contour. Dots are major mountain peaks: A = Arc Dome in the Toiyabe Range, T = Table Mountain in the Mon- itor Range. (Vegetation map derived from LANDSAT-1 photography. Elevational contours from U.S. Geologi- cal Survey maps.) 2600 . Z400 \l e f¥ r c ■ | 2200 %) 2000 VI Number of species 0 50 100 0 50 100 % of total cover % of represented by trees equitability Fig. 20. Species density (floristic richness), relative percent of total cover contributed by trees, and per- cent of maximum equitability calculated by the Mcin- tosh (1967) index plotted by elevation and aspect, Needle Range, Beaver-Iron Cos., Utah. 134 GREAT BASIN NATURALIST MEMOIRS No. 2 In general, the north slopes at any eleva- tion have a larger number of species, a lower percent relative cover of trees, and a higher equitability value than the other three aspects. Higher average diversity values were ob- tained in our data by including some of the woodland that is encroaching on shrublands and grasslands both above and below the main woodland belts. These successional considerations are sufficiently complex that details have to be considered elsewhere (Nabi 1978). matic patterns are so different in the vari- ous portions of the Great Basin. These differences are striking even at the same elevation on different portions and aspects of the same mountain range. This is espe- cially evident in data from the Shoshone Range. In general, such differences are most pronounced in the Mojave-Great Basin tran- sition area. Stratification of landscapes into units of some homogeneity for management purposes must take these phytogeographical patterns into account, if confusion is to be avoided. IMPLICATIONS Our observations of these broad phyto- geographical patterns in Great Basin juniper-pinyon woodlands will hopefully lead to some testable hypotheses about eco- physiological responses of juniper and pi- nyon. For instance, what are relative evapo- transpirative and photosynthetic limits of juniper and pinyon in regard to temper- ature and moisture? Does needle number on pinyon relate to effective soil moisture? What are the tolerances of seedlings to tem- perature and moisture extremes compared to those of the mature trees? Are there late winter-early spring low temperature suscep- tibilities? Research to answer these ques- tions provoked by our synecological results will be necessary to confirm our hunches. The distributional and phytosociological variation observed within the type indicates that effective environments are radically different across the Great Basin. For the time being we can only inferentially con- clude from the phytogeographical evidence that climate is a major factor influencing the latitudinal, longitudinal, and altitudinal extent of Great Basin juniper-pinyon wood- lands. Climate is also probably a greater de- terminant of internal floristic diversity than migrational and evolutionary equilibria. Re- garding the mountains and their woodlands as islands in a uniform "sea" of desert leads to dangerous basic conclusions and con- founds management application, since cli- ACKNOWLEDGMENTS The data were collected with financial support from the Intermountain Forest and Range Experiment Station, U.S. Forest Ser- vice. Subsequent analysis was done under support of Mclntire-Stennis Forestry Re- search Act monies through the Utah Agri- cultural Experiment Station. We wish to ac- knowledge the assistance of Robert Bayn in data analysis and preparation of figures. Literature Cited Antevs, E. 1948. Climatic changes and pre- whiteman, p. 167-191. In: The Great Basin with emphasis on Glacial and post-Glacial times. Bull. Univ. of Utah 38(2): 1-191. 1955. Geologic climatic data in the West. Amer. Antiquity 20: 317-335. Auclair, A. N., and F. G. Goff. 1971. Diversity relations of upland forests in the western Great Lakes area. Amer. Naturalist 105: 499-528. Beeson, D. W. 1974. The distribution and synecol- ogy of Great Basin pinyon-junipers. M.S. thesis, Univ. Nevada, Reno, 95 p. Billings, W. D. 1951. Vegetational-zonation in the Great Basin of western North America, p. 101-122. In: Compt. Rend, du Colloque sur Les Bases Ecologiques de la Regeneration de la veg- etation des Zones Arides. Union Intern. Soc. Biol., Paris. 1954. Temperature inversions in the pinyon- juniper zone of a Nevada mountain range. But- ler Univ. Bot. Studies 12: 112-117. Birkeland, P. W. 1969. Quaternary paleoclimatic implications of soil clay mineral distribution in a Sierra Nevada-Great Basin transect. J. Geol. 77: 289-302. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 135 Brown, J. H. 1971. Mammals on mountaintops: nonequilibrium insular biogeography. Amer. Naturalist 105: 467-478. Cottam, W. P., J. M. Tucker, and R. Drob- nick. 1959. Some clues to Great Basin post- pluvial climates provided by oak distribution. Ecology 40: 361-377. Critchfield, W. B., and G. L. Allen- baugh. 1969. The distribution of Pinaceae in and near northern Nevada. Madrono 2: 12-26. Cronquist, A., A. H. Holmgren, N. H. Holmgren, and J. L. Reveal. 1972. Intermountain flora, Vol. 1, Hafner Publ. Co., New York. Elston, R. (ed.). 1976. Holocene environmental changes. Nevada Archeological Survey Research Paper Series No. 6. Univ. Nevada Press, Reno. Fritts, H. C. 1965. Tree-ring evidence for climatic changes in western North America. Monthly Weather Rev. 93: 421-443. Hocker, H. W., Jr. 1956. Certain aspects of climate as related to the distribution of loblolly pine. Ecology 37: 824-834. Hollermann, P. 1973a. Some aspects of the geo- ecology of the Basin and Range Province (Cali- fornia section). Arctic and Alpine Res. 5(3) Part 2: A85-A98. 1973b. Some reflections on the nature of high mountains, with special reference to the western United States. Arctic and Alpine Res. 5(3) Part 2: A 149- A 160. Houghton, J. G. 1969. Characteristics of rainfall in the Great Basin. Desert Research Institute. Univ. of Nevada System, Reno. Hume, L., and J. C. Day. 1974. The determination of an efficient sampling intensity for studying beta diversity in plant communities. Can. J. Bot. 52: 189-199. Johnson, N. K. 1956. Birds of the pihon association of the Kawich Mountains, Nevada. Great Basin Naturalist 1-4: 32-33. 1975. Controls of number of bird species on montane islands in the Great Basin. Evolution 29: 545-567. Johnson, M. P., and P. H. Raven. 1970. Natural regulation of plant species diversity, p. 127-162. In: T. Dobzhansky et al. Evolutionary Biology Vol. 4. Appleton-Century-Crofts, New York. LaMarche, V. C, Jr. 1974. Paleoclimatic infer- ences from long tree ring records: Intersite comparison shows climatic anomalies that may be linked to features of the general circulation. Science 183: 1043-1048. LaMarche, V. C, and H. A. Mooney. 1972. Recent climatic change and development of the bristlecone pine (P. longaeva Bailey) krummholz zone, Mt. Washington, Nevada. Arctic and Al- pine Res. 4: 61-72. Lanner, R. M. 1976. The tree that made Nevada's silver, p. 443-444. In: S. A. Paher (ed.), Nevada Official Bicentennial Book. Nevada Publications, Las Vegas. Lowe, C. H. 1964. Arizona's natural environment: Landscapes and habitats. Univ. Arizona Press, Tucson. Lustic, L. K. 1969. Trend-surface analysis of the Basin and Range Province, and some geomor- phic implications. U.S. Geol. Surv., Prof. Pap., 500-D: 1-70. Martin, P. S. 1963. The last 10,000 years: A fossil pollen record of the American southwest. Univ. of Arizona Press, Tucson. Martin, P. S., and P. J. Mehringer, Jr. 1965. Pleistocene pollen analysis and biogeography of the Southwest, p. 433-451. In: H. E. Wright, Jr., and D. G. Frey (eds.). The Quaternary of the United States. Princeton Univ. Press, Princeton. May, R. M. 1975. Patterns of species abundance and diversity, p. 81-120. In: M. L. Cody and J. M. Diamond (eds.). Ecology and Evolution of Communities. Harvard Univ. Press, Cambridge. McIntosh, R. P. 1967. An index of diversity and its relation to certain concepts of diversity. Ecology 48: 392-404. Morrison, ^ B. 1965. Quaternary geology of the Great Basin, p. 265-285. In: H. E. Wright, Jr., and D. G. Frey (eds.). The Quaternary of the United States. Princeton Univ. Press, Princeton. Mueller-Dombois, D., and H. Ellenberg. 1974. Aims and methods of vegetation ecology. John Wiley and Sons, New York. Nabi, A. A. 1978. Variation in successional status of pinyon-juniper woodlands in the Great Basin. M.S. thesis, Utah State Univ., Logan, 127 p. Newnham, R. M. 1968. A classification of climate by principal component analysis and its relation- ship to tree species distribution. For. Sci. 14: 254-264. Phillips, A. M., and R. T. VanDevender. 1974. Late Pleistocene plant communities in the lower Grand Canyon of Arizona. Bull. Ecol. Soc. Amer. 51: 27 (abstr.). Richardson, E. A., S. D. Seeley, and D. R. Walker. 1974. A model estimating the com- pletion of rest for "Redhaven" and "Elberta" peach trees. HortScience 9: 331-332. Schnell, G. D., P. G. Risser, and J. F. Helsel. 1976. Factor analysis of tree distribution pat- terns in Oklahoma. 23rd annual meeting, South- western Assoc. Naturalists (abstr.). p. 11. Simberloff, D. S. 1974. Equilibrium theory of is- land biogeography and ecology. Ann. Rev. Ecol. and System. 5: 161-182. Spaulding, W. G. 1974. Dynamics of late Pleisto- cene vegetation change in southern Nevada. Bull. Ecol. Soc. Amer. 51: 27 (abstr.). 136 GREAT BASIN NATURALIST MEMOIRS No. 2 St. Andre, G., H. J. Mooney, and R. D. Wright. 1965. The pinyon woodland zone in the White Mountains of California. Amer. Midi. Naturalist 73: 225-239. Stebbins, G. L. 1974. Flowering plants: Evolution above the species level. Harvard Univ. Press, Cambridge. 1975. Shnibs as centers of adaptive radiation and evolution, p. 120-140. In: H. C. Stutz (ed.). Wildland Shrubs: Symposium and Workshop Proc. U.S. Forest Service, Shrub Sciences Lab, Brigham Young Univ. Press, Provo, Utah. Tueller, P. T. 1975. The natural vegetation of Ne- vada. Mentzelia 1: 3-6, 23-28. United States Department Agriculture, Soil Survey Staff. 1951. Soil Survey Manual. U.S.D.A. Agric. Res. Admin. Hdbk. 18. 1960. Soil classification: A comprehensive system, 7th Approximation. U.S. Government Printing Office, Washington, D.C. 1967. Supplement to soil classification sys- tems (7th Approximation). U.S. Government Printing Office, Washington, D.C. VanDevender, R. T. 1974. Late Wisconsin biotic communities and climates of the Sonoran Desert of Arizona: The fossil pack rat midden record. Bull. Ecol. Soc. Amer. 51: 27 (abstr.). Wells, P. V., and C. D. Jorgenson. 1964. Pleistocene wood rat middens and climatic change in Mojave Desert: A record of juniper woodlands. Science 14: 1171-1174. Wells, P. V., and R. Berger. 1967. Late Pleisto- cene history of coniferous woodland in the Mo- jave Desert. Science 155: 1640-1647. Whittaker, R. H. 1975. Communities and ecosys- tems. 2d ed. MacMillan Publ. Co., New York. Wernstedt, F. L. 1960. Climatic fluctuations in the Great Basin 1931-56. J. Range Manage. 13: 173-178. West, N. E., K. H. Rea, and R. J. Tausch. 1975. Basic synecological relationships in juni- per-pinyon woodlands, p. 41-53. In: G. F. Gif- ford and F. E. Busby (eds.). The pinyon-juniper ecosystem: A symposium. Utah Agric. Expt. Sta. Young, J. A., R. A. Evans, and P. T. Tueller. 1976. Great Basin plant communities- pristine and grazed, p. 186-215. In: R. Elston (ed.). Holocene environmental change. Nevada Archeological Survey Research Paper Series No. 6, Univ. Nevada Press, Reno. PATTERNS OF AVIAN GEOGRAPHY AND SPECIATION IN THE INTERMOUNTAIN REGION Ned K. Johnson1 Abstract— The Intermountain Region comprises a huge arena where major avifaunas more highly developed beyond the Great Basin come into contact. For this reason the distribution and composition of avifaunas in ri- parian and pinyon-juniper woodlands and in coniferous forests in this region are understood most clearly if con- sidered as part of a broader system of patterns evident in western North America. Riparian woodland samples point to a complex meeting ground in the Intermountain Region of two major avifaunas of equivalent size, but from opposite distributional backgrounds. These northern and southern avifaunas do not mix among the samples represented except in western Nevada where two species of ultimate southern origin have penetrated a basically northern avifauna. Approaching the Great Basin, from the two centers of abundance of riparian species in the Snake and Colorado River drainages, species richness drops. Habitat depletion and, to a lesser extent, insularity play roles in this impoverishment. In the most depauperate riparian avifaunas, six species commonly coexist, each in a different family. Comparison of the pinyon zone avifaunas of two groups of mountain ranges, 90 km apart along the California-Nevada border, demonstrates a striking trade off among species of northern and southern biogeographic histories. The northern or Boreal forms, many of which are numerous in the Sierra Nevada, have had easy access to the favorable, cool, and relatively moist pinyon forests in the adjacent spur ranges. In contrast, the species of southern or Austral derivation prefer the warm and very arid pinyon woodland a short distance to the south. Few species are confined to either northern or southern sites, overlap in species composition is great, and equivalent species richness is achieved. However, despite these similarities, strong geographic differences in abundance of most species in each pinyon avifauna and the occurrence of at least 12 specific and subspecific range boundaries suggest the interposition between northern and southern pinyon areas of a substantial, but as yet poorly characterized, climatic barrier. Boreal species richness declines,abmptly from high values in the south- ern Cascades and Sierra Nevada to low levels in a zone of impoverishment across western and central Nevada. From near 116° W Longitude in eastern Nevada, coincident with the appearance of fir and/or bristlecone pine forests, species numbers climb gradually until the main Rocky Mountains are reached. There, species richness compares favorably with that in the Sierra Nevada. The proportion of species favoring riparian woodlands over coniferous forests is higher on the island mountaintops of the Great Basin than in the Boreal "continents" of the Sierra Nevada-Cascades and Rocky Mountains. The western edge of the Great Basin richly demonstrates exam- ples of stages in avian speciation. A full range of interactions is represented, from intergradation of poorly char- acterized races, through abrupt zones of hybridization between strongly marked subspecies of different racial complexes or of semispecies, to sympatry and infrequent interbreeding of closely-related, full species in recent secondary contact. These various zones of population interaction coincide strikingly with sharp floristic and cli- matic gradients. A major avifaunal break occurs between coastal or Sierra Nevadan forms that inhabit oak- chaparral and/or coniferous forest and closely-related interior forms that prefer pinyon-juniper or aspen-willow associations. In keeping with the special requirements of each species, contact zones and areas of disjunction show general, rather than precise, coincidence in the western Great Basin. There is no Great Basin Boreal Avi- fauna; the most distinctive interior forms occur across the entire span from eastern California to Colorado. The- low desert trough along the east side of the Sierra Nevada that divides major mountain systems in the western portion of the Intermountain Region is not the principal barrier dividing coastal Sierra Nevada from interior avi- faunas. Instead, the major avifaunal transition occurs in a belt of variable width just east of the crest of the Cas- cade Mountains and Sierra Nevada, where the precipitation shadow and continental climate begin to exert cru- cial influence. Although much neglected in the past be- mountain Region of western North America cause of difficulty of access and false im- is emerging as a fascinating natural labora- pressions of biotic sterility, the Inter- tory requiring the close attention of the 'Museum of Vertebrate Zoology and Department of Zoology, University of California, Berkeley, California 94720. 137 138 GREAT BASIN NATURALIST MEMOIRS No. 2 biogeographer-ecologist. Islands of conifers or aspen, cool mountain meadows, and at- tendant biotas on scores of mountaintops lie isolated by seas of desert from source stocks in the Sierra Nevada and Rocky Mountains. The unique physical setting of a multi- plicity of more or less parallel ranges, each differing in size, geology, and configuration from adjacent ones, cannot be duplicated elsewhere among continental mountain sys- tems. And, because of low human density, many of the montane habitats are relatively undisturbed, thus permitting a good look at original, man-unmodified distributions of species. In the valleys as well, oases offer insular environments for a variety of organ- isms whose interrupted ranges intrigue the biogeographer. With increased field efforts by a number of workers in recent decades, the bird spe- cies breeding on many of the major basin ranges have been tallied, and information has reached a level allowing tests of certain fundamental theories of insular biogeog- raphy (Johnson 1975). However, these tests are just the beginning. With additional "al- pha" exploration in the future, the enlarged data base will permit more diverse and de- tailed kinds of analyses. But, already enough distributional information has accumulated to suggest several patterns that challenge the avian geographer. In this paper I identi- fy and discuss these patterns in a search for common biogeographic themes. Further- more, I specify among the birds of the In- termountain Region distributional situations relevant to hypotheses on the control of range boundaries. Finally I describe the prominent zone of active speciation in the western part of the region and attempt to interpret this zone in relation to coincident physiographic, climatic, and floral dis- continuities. Nomenclature and sequence of species follow the checklist of North American birds except where superceded by the 32nd supplement (American Ornithologists' Union 1957, 1973). Avifaunal Transects in the Intermountain Region Riparian woodlands in valleys and can- yons, pinyon woodlands on mountainsides, and coniferous forests and aspen on moun- taintops comprise the major habitats for birds in the region between the axis of the Sierra Nevada-Cascade Range and the Rocky Mountains. Analyses of the richness and composition of selected avifaunas breeding in sample areas of these woodlands and forests expose several patterns that elucidate the fundamental organization of avian distribution in the region. Moreover, these samples along transects demonstrate the complex modes by which historical as- pects of avifaunal distribution influence present-day patterns of species richness and community structure. Figure 1 shows the lo- cation of the sample areas and their basic avifaunal relationships. Latitudinal Variation in Riparian Woodland Avifaunas Bird species requiring riparian vegetation for breeding occur discontinuously in cot- tonwoods, willows, ashes, and associated thickets along watercourses and at valley oases throughout the Great Basin region. When species of land birds breeding in such habitats are compared (Table 1), among 11 sample areas from the Snake River drainage in the north to the Colorado River drainage in the south, prominent differences and some unexpected similarities emerge in total species numbers, composition of the avi- faunas according to evolutionary derivation, and structure of the avifaunas by ecologic roles. That these three topics are inter- related will become evident in the dis- cussion to follow. Total species numbers.— Of immediate in- terest is the striking similarity of the geo- graphically extreme stations, Central Idaho and Colorado River, in species richness, 28 versus 25, from the pool of 43 species. But, upon entering the Great Basin from the Snake River drainage, species totals drop, to 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 139 21 in the Elko sample, 17 in Humboldt, 15 in Toiyabe, and to 12 in Kawich. Similarly, moving north from the Colorado River, to- tal species numbers fall to 13 in Pahranagat, 12 in Meadow Valley Wash, and 11 at Ash Meadows-Pahrump. Figure 2 demonstrates the strong and abrupt avifaunal attenuation that occurs with increasing distance from the lush riparian woodlands to the north and south. The smallest tallies, for the Kaw- ich area and Ash Meadows-Pahrump, prob- ably reflect habitat impoverishment in view of the limited extent and interrupted growth of the riparian clumps there. For Fig. 1. Sample areas of riparian woodland (squares and circles) and pinyon woodland (triangles) avifaunas in the Intermountain Region. Shading grossly defines regions of forest and woodland. Numbers in squares or circles correspond to those of sample areas in Table 1. Arrows indicate closest avifaunal relationships. Northern and southern pinyon areas are denoted by solid and open triangles, respectively. Approximate boundary of northern and southern riparian woodland avifaunas is shown by the dashed line between samples 7 and 8. 140 GREAT BASIN NATURALIST MEMOIRS No. 2 Table 1. Species breeding in riparian woodlands at selected localities in the Intermountain Region'. 123456789 10 11 s£ c3 a- Northern Element Coccyzus erythropthalmus X Dendrocopos villosus2 X X X X X X X - - - - Dendwcopos pubescens2 X X X X1 - X - - - - - Tyrannus tyrannus X - X — ' Iridoprocne bicolor X X - X - X - - - - - Parus atricapillus1 X X Troglodytes aedon X X X X X X X - - - - DumetelLi carolinensis X Turdus migratorius X X X X X X X - - - - Catharus fuscescens X X - - - X - - - - - Catharus ustulatus X X X X X X - - - - - Vireo gilvus X X X X X X X - - - - Vireo olivaceus X Vermivora celata X X X - X X - - - - - Oporornis tolmiei X X X X X X X - - - - Stetopliaga ruticilla X Passerella iliaca X X X X X X - - - - - Southern Element Zenaida asiatica -' -' X Dendrocopos scalaris2 X X - X Centurus uropygialis' X Sayomis nigricans X X - X Pyrocephalus rubinus2 X - X X Thryomanes bewickii2 - - - X' - - - X X X X Toxostoma dorsale2 X X Sialia mexicana - - - X1 - - - - - - - Phainopepla nitens2' -5 -' X Vireo bellii x< X Vermivora luciae -5 X - X Icterus cucullatus X X X Piranga rubra -5 X Guiraca caerulea -5 X X X X Pipilo aberti X - X Widespread Element Falco sparverius2 X X X X X X X X X X X Coccyzus americanus X - - X - - - X - X X Otus asio2 X - - X - - - X X - X Archilochus alexandri X X - X - X - X - - X Colaptes auratus2 X X X X X X X - X - X Tyrannus verticalis X X X X X X X X X X X Empidonax traillii X X -5 X -5 - - - - X X Dendroica petechia X X X X X X X - - - X Icteria virens X X X X X X X X - - X Icterus galbula X X X X X X X X X X X Melospiza melodia2 X X X X' X X X X - - X Totals 28 21 17 22 15 19 12 13 12 11 25 'Sources of data on occurrence as follows: (1) Central Idaho, a composite list from several localities in the center of the state, taken from Burleigh (1972); (2) Elko, a composite list from field notes and specimens in the Museum of Vertebrate Zoology ( = MVZ) from the northern portion of Elko County, and from Linsdale (1936); (3) Humboldt, from Tavlor (1912); (4) Truckee-Carson. a composite list from original data of author (= NKJ) and from Linsdale (1936); (5) Toivabe, from Linsdale (1938); (6) Ely, data from Spring and Steptoe valleys, NKJ; (7) Kaw.ch, NKJ; (8) Pahranagat, NKJ and MVZ; (9) Meadow Valley Wash. NKJ and Linsdale (1936); (10) Ash Meadows-Pahrump; (11) Colorado River, NKJ and Grinnell and Miller (1944). Spe- cies that typically prefer or require riparian woodland are included. 'Permanently-resident species. 'Form shows closest affinity to populations in California, to the northwest. 'Former or casual occurrence. 'Recorded during the breeding period, but not proved to be nesting. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 141 Ash Meadows-Pahrump and Pahranagat samples, we also may be seeing the effects of insularity. Several Colorado River drain- age species, such as the Abert's Towhee (a highly sedentary, permanently resident form), probably have had great difficulty colonizing the isolated riparian areas to the north. However, one could also argue that this and other potential colonists are missing from these remote areas because the proper kinds of riparian habitats, as found along the Colorado River and its tributaries, are lacking. The moderate size (22 species) of the Truckee-Carson sample reflects its position at the well-watered eastern base of the Car- son Range. Origin of riparian woodland avifaunas.— The entire riparian woodland avifauna can be divided on the basis of evolutionary D widespread ED northern £/] southern I. Central Idaho 39% 50 43 42 47 58 45 42 : 6? 44% 4 Truckee-Carson 2 Elko 6 Ely 3 Humboldt 5 Toiyabe 7 Kawich 10 Ash Meadows-Pahrump 9 Meadow Valley Wash 8 Pahranagat II. Colorado River 5 15 25 Density and origin of riparian woodland bird species Fig. 2. Percent composition, according to geographic derivation, of avifaunas in 11 riparian woodland sample areas. These are arranged in order of decreas- ing avifaunal size (species density) toward the bound- ary of northern and southern avifaunas, between Ka- wich and Ash Meadows-Pahrump. For example, the Colorado River sample has 25 species, 11 (= 44 per- cent of which are widespread and 14 ( = 56 percent) of southern derivation. background or derivation into (1) a northern element consisting of species with northern centers of distribution and affinities, (2) a southern element comprised of species with southern distributional relationships, and (3) a widespread element including species oc- curring throughout the western United States or beyond. The species are grouped in Table 1 according to these three ele- ments. Note that none of the species of the northern element occurs in any samples south of Kawich and, with one exception (see below), none of the species of the southern element is found in any samples north of Meadow Valley Wash or Pahrana- gat. Therefore, the two major avifaunas ap- proach, but do not contact and intermix, in the southern Intermountain Region. Although basically of northern derivation, the riparian avifauna of the Truckee-Carson sample is of special interest because it con- tains a distinctive group of four species with affinities to populations of northeastern California. One of these species, the Downy Woodpecker, is of northern affinity; two species, the Western Bluebird and Bewick's Wren, are of southern derivation. The Song Sparrow is widely distributed. The western Nevadan populations of these four species probably originated from stocks in the Cen- tral Valley of California, stocks that spread into the western Great Basin via the Modoc Plateau. The widespread component comprises be- tween 39 and 62 percent of each sample (Fig. 2). No clear evidence of differential decline of northern (or southern) versus widespread elements is shown as the zone of impoverishment is approached. Thus, the diminution is general and involves approx- imately equivalent shrinkage of the two ma- jor components of each avifauna. Community structure and ecologic roles.— A surprising finding is that the widespread species fraction seems to have stabilized, at either seven or five species, for the five sample areas with the smallest numbers of species. Thus, the Humboldt, Toiyabe, and Kawich samples have seven widespread spe- 142 GREAT BASIN NATURALIST MEMOIRS No. 2 cies in addition to northern ones. And, the lists of species are identical! The Ash Meadows-Pahrump and Meadow Valley Wash samples include southern species and five that are widespread. However, in these two samples the coincidence in identity of widespread forms is incomplete; each shares three of their five species. Six of the 11 widespread species, Falco sparverius, Colaptes auratus, Tyrannus verti- cals, Icteria cirens, Icterus galbula, and Melospiza melodia, occur in nine or more sample areas. Because these species com- monly coexist, they form a group of great interest. I view them as a "standard riparian woodland species" group analogous to the "standard Boreal species" group I identified in montane habitats of the western United States (Johnson 1975). Note that each of the six species is in a different family. There- fore, coexistence in the riparian woodland avifaunas examined here predominates among species of fundamentally different body designs. Included in the tallies of Table 1 are five pairs of closely related species, three pairs of which are congeneric. Furthermore, within each pair, each species is from a dif- ferent distributional element. The pairs are Coccyzus erythropthalmus, and C. american- us, Dendrocopos pubescens and D. scalaris, Tyrannus tyrannus and T. verticalis, Troglo- dytes aedon and Thryomunes bewickii, and Dendroica petechia and Vennivora luciae. Not all of the species of each pair contact locally, but the interactions of those that do meet would merit close study. Indeed, a fertile field awaits the ecomor- phologist willing to analyze community structure, foraging roles, and morphologic adaptations among the coexisting species of birds in any of the riparian woodlands dis- cussed. Latitudinal Variation in Pinyon Woodland Avifaunas Among arboreal habitats, groves of single- leaf pinyon (Pinus monophylla) comprise the most extensive formation in the Inter- mountain Begion, and these woodlands therefore represent an important habitat for breeding birds. Typically this pine is scat- tered over mountain slopes in open stands or clumps of small trees with much inter- vening brush. Occasionally, on favorable sites, growth can be luxuriant, and true for- est is achieved. Here trees grow to heights of 10 m or more, and the closed canopy shades a needle-strewn forest floor. This site-to-site variability in growth form of pi- nyon intrigues the ornithologist because, as has long been known, birds clearly respond to physical features of vegetation. A second noteworthy aspect of pinyon is the wide elevational range often occupied on a single mountainside. In the White Mountains, Cal- ifornia, for example, pinyon occurs from ap- proximately 2000 to 2900 m, with the best development between 2450 and 2600 m (St. Andre et al. 1965). In the Grapevine Mountains of southern Nye County, Nevada, these variable charac- teristics of pinyon are seen clearly (Miller 1946). Small, scrubby trees grow in scat- tered stands, exceptionally fine tracts of old trees with trunks two feet in diameter form forests locally on the high northeast slopes, and the pinyon zone is wide, extending to 2650 m at the top of the highest peak. The breeding avifauna has responded to these differences. The complement of species that normally lives in pinyon of usual form and spacing is present, and several kinds of birds that typically avoid this plant formation breed in the most luxuriant stands. Evi- dently the latter species are attracted to physiognomic features of the pinyons that resemble those of the coniferous forest they ordinarily inhabit. Furthermore, at least three additional species that otherwise would be absent occur in the cool upper reaches of the zone, apparently responding there to preferred or required summer tem- perature regimes. I also have noted similar occurrences of birds in unusually dominant pinyon woodland in the Kawich Mountains of central Nye County, 100 km north- northeast of the Grapevines (Johnson 1956). 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 143 In recent years I have explored additional areas of heavy pinyon that invite avifaunal comparison with the groves in the Grape- vines. These stands of old trees grow local- ly, in Nevada near the California line, in the Palmetto Mountains ( = Mount Ma- gruder plus the Silver Peak Mountains), Es- meralda County; the Wassuk Range, Miner- al County; the Pine Grove Hills, Lyon County; and in the Sweetwater Mountains, Douglas and Lyon counties. In each of these regions, open woodlands of small trees and stands of intermediate height and den- sity predominate on most slopes; the forests are more local in distribution. For purposes of comparison of the avi- faunas, it is useful to consider these several pinyon areas as comprising (1) a northern subset occupying the spur ranges connected to the east side of the Sierra Nevada, which includes the Wassuk and Sweetwater moun- tains and the Pine Grove Hills, and (2) a southern subset of wooded islands lying to the southeast, including the Palmetto and Grapevine mountains. The massive White Mountains lie between the two groups. The northern and southern subsets (Fig. 1) per- mit comparison of avifaunas along a north- west to southeast gradient. Some striking differences in avifaunal composition and nu- merical relationships of species emerge (Table 2). Although the northern and south- ern localities are only 90 km apart, several species reach their distributional limits be- tween these mountain systems and thus breed in only one of the two regions. Unex- pectedly, 19 species change in average pop- ulation density along the gradient. For some (e.g., Empidonax ivrightii), the change is subtle. For others (e.g., Vireo solitarius) population densities are dramatically differ- ent between northern and southern moun- tains. Several species partly or completely switch habitats. For example, Thryomanes bewickii and Otus asio, common and un- common residents, respectively, in the pi- nyon woodlands of the southern Great Basin and northern Mojave Desert, prefer riparian cottonwoods and willow in west-central Ne- vada. Although both species use pinyon at the more northerly sites, the wren is un- common, and the owl is rare in this associ- ation. Empidonax wrightii nests in tall sagebrush, bitterbrush, and juniper in the northwestern Great Basin where pinyon is lacking but is scarce or absent in such vege- tation in southern Nevada. In other species the habitat change is more subtle. Parus in- ornatus and Dendroica nigrescens inhabit pi- nyon in both northern and southern local- ities, but their numbers diminish in the north where they prefer the more arid and warmer portions of the pinyon belt. In 5 of the 19 species, the change in density be- tween north and south coincides with a shift in subspecific status and, in 3 species, with a concurrent change in habitat prefer- ence (Table 2). How are we to interpret these contrasts in the avifaunas of the pinyon association of closely adjacent geographic regions? Two interrelated issues, (1) variation in the cli- matic aspects of the pinyon zone environ- ment and (2) temperature-moisture prefer- enda and distributional histories of the bird Table 2. Species composition and numerical rela- tionships of pinyon woodland avifaunas of spur ranges of the northern Sierra Nevada1 versus those of mon- tane islands to the southeast2. Boreal (B) and Austral (A) species are distinguished. Species present only in north: Glaucidium gnomu3(B) Sphyraupius ruber3 (B) Cyarwcitta stelleri (B) Species present in both areas but Aegolius acadicus (B) Dendrocopos villosus4 (B) Tachycineta thalassina (B) Parus gambei45 (B) Sirto carolinensis (B) Species present in both areas but Otus asio*5 (A) Empidonax wrightii5 (A) Aphelocoma coerulescens (A) Psaltripurus minimus4 (A) Species present only in south: Regulus calendula3 (B) Vireo vicinior (A) Certhia familiaris3 (B) Myadestes townsendi (B) more numerous in north: Sialia currucoides (B) Piranga ludoviciana (B) Carpodacus cassinii (B) Chlorura chlorura (B) Junco huemalis (B) more numerous in south: Thyromanes bewickii45 (A) Polioptila caerulea5 (A) Vireo soliarius (B) Icterus parisorum (A) Junco caniceps (B) 'Wassuk Range, Sweetwater Mountains, and Pine Grove Hills. 2Grapevine Mountains and Palmetto Mountains (= Mt. Magruder + Silver Peak Mountains). 3Breeds onlv verv locally in pinyon woodland in the northern or south- em mountains. 4Subspecies in north different from that in south 5Partial or complete habitat shift from north to south. 144 GREAT BASIN NATURALIST MEMOIRS No. 2 species at northern and southern stations, may offer partial explanation. From rainfall data for the nearby White Mountains and several other pinyon sites in the southwest, St. Andre et al. (1965) con- cluded that overall aridity characterizes this plant formation. Unfortunately, no moisture data exist for the pinyon zones in the ranges of special interest here. Nonetheless, pinyon certainly occupies a range of rainfall regimes, and the more northern sites, al- though still relatively arid, presumably have greater average rainfall than the southern sites. Temperature data also are scarce, ex- cept for the White Mountains. However, I can conclude at least that the upper part of the zone averages considerably cooler than the lower part, a fact readily apparent to the researcher working at different eleva- tions in this region. The unusual upper ex- tension of the pinyon zone in the White Mountains (St. Andre et al. 1965) and in the Grapevine Mountains (Miller 1946) would enhance temperature differences that nor- mally exist between the different elevations. Finally, because of the difference in lati- tude, it is probable that the northern sites average cooler than the southern ones. The climatic preferences and distribution- al backgrounds of the bird species agree closely with the groupings of Table 2. The 5 species confined to the north and all 10 forms that are more numerous in the north are Boreal species with distributional back- grounds associated with northern coniferous forests. In contrast, except for Regulus cal- endula, Vireo solitarius, and Junco caniceps, forms of Boreal derivation (Miller 1951), all the species present only in the south or reaching their greatest numbers in the southern pinyon areas are of Austral origin, having been associated in their distribution- al histories with southwestern pinyon or oak woodlands. I conclude that this major trend in composition and density of pinyon zone avifaunas along the Nevada-California bor- der apparently results from the exchange along a temperature-moisture gradient of southern pinyon woodland species, adapted principally to the warm and relatively arid portions of the zone, and northern species, adapted chiefly to the relatively cool and moist portions of the zone. The origin of the bird species of the Wassuks, Sweetwaters, and Pine Grove Hills is of particular interest because the pinyon woods in these ranges directly contact the coniferous forests of the Sierra Nevada. For this reason, bird populations of the pine-fir zone of the east side of the latter range have had easy access to pinyon forests lying to the east. Ready access may be one reason why Cyanocitta stelleri and, locally, Certhia familiaris and Sphyrapicus ruber breed in pinyon in these areas. Otherwise these spe- cies are not known to nest in this forma- tion. It is noteworthy that none of these forms inhabits the old pinyon groves in the Desatoya Mountains of central Nevada, pre- sumably because of the great distance of this range from source populations. Note that from the pool of 27 pinyon zone species, nearly identical total numbers of breeding forms, 23 and 24, respectively, are recorded for northern and southern sta- tions. Such equivalence is significant. Im- portantly, no neat system of replacement or of switches in abundance by species with similar ecologic roles occurs along the gradient (Table 2). Instead, the evolution of community structure at various localities in the pinyon formation probably has involved more subtle adjustments among the species in foraging zones, food quality, and /or habi- tat preference. Those species of birds repre- sented by different geographic races at northern and southern localities would be especially worthy of close study in this re- gard. Gradients in Boreal Avifaunas I recently discussed controls of richness of Boreal species on 11 sample areas in the Sierra Nevada, Cascades, and Rocky Moun- tains and on 20 mountaintop islands in the Intermountain Region (Johnson 1975). Here I wish to expand on portions of that analy- sis through examination of both longitudinal and latitudinal trends in numbers of species 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 145 occupying particular Boreal habitats along- four transects across the Great Basin. These transects (Fig. 3, lines I-IV) are roughly equidistant latitudinally and pass through 20 of the sample areas described in my pre- vious paper. Species numbers now have been determined for eight additional areas, and these new data are included. Transect HI fails to cross the entire Intermountain Region because no species lists are available for restricted portions of the main Rocky Mountains directly east of the Snake Range in eastern Nevada. Total species richness.— All transects demonstrate the abrupt drop in total num- bers of species in the north-south-oriented swath that extends across much of western and central Nevada (Fig. 4A). From high species totals of 56 to 64 at Mt. Shasta, Lassen, Tahoe, and Yosemite, species rich- ness declines sharply along transect I, be- tween Warner and Pine Forest; along trans- ect II, between Lassen and West Humboldt; along transect III, between Carson and Pine Nut; and along transect IV, between Yose- mite and Glass Mountain. Species numbers then climb toward the east to totals of 56 and 55 at Uinta and Aquarius-Boulder, re- spectively, at the ends of the tran- sects. Highest totals for the Sierra Nevada- Cascades always equal or exceed those for the Rocky Mountains. In the region of gen- eral impoverishment along each transect, a mountain range moderately rich in Boreal species interrupts the general smoothness of each species density curve (Fig. 4A): I, Jar- bidge; II, Ruby; III, Toiyabe-Shoshone; and IV, White-Invo. Fig. 3. Four transects across the Intermountain Region, passing through 28 sample areas of Boreal avifaunas. These sample areas are named in Fig. 4. No native firs are known to occur between the two vertical dashed lines. 146 GREAT BASIN NATURALIST MEMOIRS No. 2 Species richness according to habitat.— In Figure 4B I portray the percentage of Bo- real species in each of four major habitats (Johnson 1975: 549), i.e., Aquatic, Wet Meadow, Riparian Woodland, and Con- iferous Forest. This permits study of pos- sible changes along gradients of proportions of species preferring these habitats. For each transect I have calculated the average percentage of species in each of the four habitats for Sierra Nevada-Cascades sam- ples, Great Basin Islands samples, and Rocky Mountains samples (Table 3). Between-transect comparisons of propor- tions of species in continental samples in- dicate striking similarities. For example, rz II II I If " a> S>i2 2s o o £ .£= o -c oIq.q.1 t- en ♦-Aquatic Wet Meadow • "^—Riparian Woodland .*^ Coniferous Forest A. Boreal bird species richness B. Composition of boreal avifaunas Fig. 4. Numbers of Boreal bird species (left) and percent avifaunal composition by four major habitats (right) along four transects across the Intermountain Region. Continental sample areas are underlined; island sample areas are not. Twenty sample areas are described more hilly in Johnson (1975). Species totals for eight new ones added are as follows: A, Mt. Shasta (Grinnell and Miller 1944); I, West Humboldt Range (author's unpubl. data); L. Wasatch (Behle and Perry 1975); M, Tahoe (Orr and Moffitt 1971); O, Pine Nut Mtns. (author's unpubl. data); R, White Pine Mtns. (author's unpubl. data); S, Schell Creek-Egan Ranges (author's unpubl. data); and Y, Pahute Mesa (Hayward et al. 1963). 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 147 number of species in coniferous forests var- ies only from 62.5 to 69.6 percent of the to- tal Boreal avifauna in the Sierra Nevada- Cascades, and those of riparian woodland vary only from 23.2 to 24.6 percent in the several samples from the Rocky Mountains. Other habitats on continents illustrate sim- ilar uniformity. Island samples are more complex and show differences among trans- ects. Specifically, the island samples of transects I and II have proportionately more aquatic and riparian woodland spe- cies, but fewer coniferous forest species, than those of transects III and IV. Within-transect comparisons (transects I, II, and III) point to more riparian woodland species and fewer coniferous forest species in samples for the Great Basin Islands ver- sus the Sierra Nevada or the Rocky Moun- tains. However, the ecologic preferences of birds in island samples resemble those of birds in Rocky Mountains samples more than those of forms in the Sierra Nevada- Cascades. The avifaunas of transect IV do not follow the pattern of the first three ir. that the proportions of species in the sever- al major habitats are similar among samples of the Sierra Nevada, Great Basin, and Rocky Mountains. The composition of coniferous forest avi- faunas.— The fraction of coniferous forest bird species deserves special comment be- cause some of the trends in total species richness discussed earlier are explained part- ly by the distribution of suitable tree spe- cies. The following group of forest bird spe- cies, distributed discontinuously in the Intermountain Region, illustrates this point: Sphyrapicus varius, Sphyrapicus thyroideus, Empidonax hatnmondii, Empidonax difficilis, Table 3. Average percent composition, according to habitat preferences, of species in avifaunas along trans- ects. Transect I Aquatic- Wet Meadow Riparian Woodland Coniferous Forest Transect II Aquatic Wet Meadow Riparian Woodland Coniferous Forest Transect HI Aquatic Wet Meadow Riparian Woodland Coniferous Forest Transect IV Aquatic- Wet Meadow Riparian Woodland Coniferous Forest Sierra Rocky vada-Cascades Great Basin Mountains Continent Islands Continent (Mt. Shasta) (Warner through Raft River) (Uinta) 3.6 3.7 5.3 12.5 10.8 16.1 14.3 34.0 23.2 69.6 51.5 (W Humboldt 55.4 (Lassen) through (Wasatch and Spruce-S. Pequop) Uinta) 7.8 6.3 3.6 12.5 4.0 17.3 17.2 37.1 24.6 62.5 52.6 (Pine Nut 54.6 (Tahoe and through Carson) Snake) 4.4 0.9 - 14.8 10.8 - 16.5 27.4 - 64.4 60.9 - (Yosemite (White-Inyo and Glass through (Aquarius- Mountain) Highland) Boulder) 2.3 0.6 3.7 12.3 7.0 14.5 19.0 23.4 23.6 66.4 69.0 58.2 148 GREAT BASIN NATURALIST MEMOIRS No. 2 Sitta canadensis, Certhia familiaris, Regains satrapa, Regulus calendula, Hesperiphona vespertina, and Spinas pinus. Although most of these forms occasionally breed in other kinds of conifers or in deciduous trees, they prefer firs (Abies concolor or Pseadotsaga menziesii) and avoid or nest sparingly in forests lacking these trees. Thus, the distri- bution of firs in the Intermountain Region is of special significance for these species. Figure 4A indicates the presence of firs in the various sample areas. Note the repeated instances of increased numbers of species, of all kinds of birds and of coniferous forest birds, in relation to the appearance of firs. Firs are absent from the zone of avifaunal impoverishment across western and central Nevada (Fig. 3). Although skimpy clumps grow in the Sweetwater Mountains, no firs occur in any of the other ranges along the California-Nevada border southeast of the Carson Range. Firs reappear in eastern Ne- vada between 115° and 116° W Longitude. Among the ranges considered here, sub- stantial stands occur in the Jarbidge Moun- tains, White Pine Mountains, and on Mount Irish. Critchfield and Allenbaugh (1969) re- port white fir locally in the Ruby Moun- tains, but the avifauna at that site has not been explored. Rristlecone pine (Pinus longaeva) domi- nates the upper slopes of several ranges in eastern Nevada. This tree is important for coniferous forest birds in part because it oc- casionally forms closed, well-shaded stands with strong physiognomic resemblance to firs. Apparently in response to this feature, several kinds of birds from the above list of- ten occupy well-developed bristlecone pine forests when firs are scarce or lacking. The two species of Sphyrapicus, Empidonax dif- ficilis, Sitta canadensis, and Regains calen- dula exemplify this point. In Nevada, the best stands of this conifer grow east of the right vertical dashed line on Figure 3 that marks the boundary of white fir distribu- tion, and in many places the two species occur in mixed stands. Breeding Range Boundaries and Interspecific Distributional Relationships Broad trends in avian geography are best identified through comparison of major components of avifaunas. Other, more de- tailed issues can be seen most clearly by in- spection of the distributions of single spe- cies and their close relatives. Here I discuss four contrasting examples, chosen from a long list of possible cases, that illustrate particular species-level distributional phe- nomena occurring among birds of the Inter- mountain Region. Three of these examples also support the view offered earlier that this region presently serves as an important arena of disjunction, contact, and inter- mixture of species representing avifaunas of differing geographic derivation. Grace's Warbler and ponderosa pine.— In southeastern Nevada, as in the southern Rocky Mountains generally, Grace's War- bler (Dendroica graciae) breeds commonly and exclusively in forests of ponderosa pine (Pinus ponderosa). But near the north- western edge of the occurrence of the Rocky Mountain form of the ponderosa pine (P. ponderosa var. scopulorum), for ex- ample in the Quinn Canyon Range, the scattered and stunted trees occupy marginal sites, and the bird is rare. It is absent from the few ponderosas that grow locally in the Highland Range, on Mount Wilson, and in the southern White Pine Mountains, and, farther north, from the extensive ponderosa pine habitat in the Snake Range (Fig. 5). This case is particularly interesting be- cause prior to 1963 the Grace's Warbler was not known to breed anywhere in south- ern Nevada. Since then, numerous records have been obtained, several from places that previously had been well-explored (Johnson 1965, 1973, 1974). Therefore, the evidence points to an active northwestward range expansion in the past decade or two. Although further extension into presently unoccupied ponderosa pine habitat is pos- sible, total distributional evidence suggests 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 149 that the approximate northwestern limit in the Intermountain Region already has been reached. However, given continued range expansion, eventual colonization of the arid ponderosa pine forests of southern Califor- nia seems probable, for the species is al- ready vagrant to that area, and a number of other species of the interior Southwest have become established there in recent years (Johnson and Garrett 1974). Insofar as is known, Grace's Warbler has never nested in the vast ponderosa pine forests of the cen- tral and northern Rockies, or in the Sierra Nevada. The warbler thus illustrates two bio- geographic phenomena typically seen in species with well-documented distributions: (1) the actual range falls short of the geo- graphic extent of seemingly appropriate habitat and (2) the distributional boundaries are unstable because of their complex con- A Grace's Warbler v 0 Ponderosa Pine "• Fig. 5. Approximate breeding distribution of Grace's Warbler (triangles) and occurrence of ponderosa pine (stippled areas) in southeastern Nevada and south- western Utah. Lower elevational perimeters of wood- lands in mountains also are indicated. trol by a multiplicity of interacting factors that are themselves fluctuating. I wish to emphasize that no other species of warbler or other pine-foliage insectivore obviously competes with D. graciae and, for this rea- son, has adjusted its distribution and abun- dance as a result of recent population inter- action with that species. Instead, the northwestern distributional perimeter of Grace's Warbler seems to be set by other aspects of the environment. For example, this warbler may be dependent upon popu- lations of pine arthropods that fluctuate in annual density, especially near the per- iphery of their ranges. Or, summer temper- ature and moisture barriers dictated by the metabolic requirements and preferenda of the warbler, the arthropods upon which it feeds, or the pines they both inhabit, may control the range border. I have no data to support a convincing explanation. Downy Woodpecker and Ladder-backed Woodpecker in disjunct allopatry.— These two congeners fail to contact in the Great Basin. The Downy inhabits aspen groves in the northern and eastern mountains (the form D. p. leucurus), or cottonwoods and willows along rivers in west-central Nevada (the form D. p. turati), and the Ladder-back lives in the cottonwood-willow association in valleys or Joshua tree woodland in the Mojave Desert. Unexpectedly, substantial areas of apparently suitable habitat occur in the wide zone now unoccupied by either species. Thus, the Downy Woodpecker has not been found in the fine aspen woodland in the Schell Creek Range or on Mount Wilson, among a number of seemingly ap- propriate places. Similarly, the Ladder-back avoids considerable Joshua tree growth in Esmeralda and Nye Counties, Nevada. Fi- nally, neither species breeds in the excellent riparian woods in Owens Valley, although the Ladder-back has been recorded at Lone Pine, at the foot of the valley (Grinnell and Miller 1944). I conclude that the respective southern and northern limits of these two species, within suitable habitats, are dic- tated primarily by temperature and mois- 150 GREAT BASIN NATURALIST MEMOIRS No. 2 ture. Considering the wide unoccupied swath between their ranges in the central Intermountain Region, distributional limits there certainly are not controlled by inter- specific competition. Mountain Bluebird and Western Bluebird in parapatry.— These two congeneric- thrushes (Sialia currucoides and S. mexicana, respectively) breed in contiguous allopatry, or parapatry, in much of the southern Inter- mountain Region. One might conclude that such a distributional relationship is the re- sult of interspecific competition (Fig. 6). In at least three general areas where both spe- cies have been recorded (Panamint Range, Delamar Range, Zion National Park), their spatial relationships are unclear. But, be- cause of their well-known differences in habitat preference (Grinnell and Miller 1944), competitive interaction probably is infrequent or absent, given a situation of lo- cal sympatry. Until more details are avail- able, I conclude that the striking parapatry results from the geographic position of a sharp shift in average spring and summer temperatures. The approximate coincidence of the zone of parapatry with the northern limits of the Mojave Desert supports this suggestion. Hammond's Flycatcher and Western Fly- catcher in broad sympatry.— As in the pre- vious examples, this case also concerns a species of northern derivation, Hammond's Flycatcher (Empidonax hammondii), and a congener of southern derivation, the West- ern Flycatcher (E. difficilis). These two forms offer a particularly impressive ex- ample of how the complex interplay of phylogenetic and distributional history has influenced present-day habitat selection and ecologic interaction. Figure 7 shows the dis- tributional and numerical relationships of these species in western North America. Hammond's Flycatcher is commonest in cool, moist Boreal forests of the Sierra Ne- vada, Cascades, and Rocky Mountains; its principal range therefore surrounds the In- termountain Region on three sides. In con- A Mountain Bluebird A Western Bluebird Fig. 6. Breeding distributional relationship of the Mountain Bluebird {Sialia currucoides) and Western Bluebird (S. mexicana bairdi) in southern Nevada and adjacent states. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 151 trast, the Western Flycatcher reaches its Mountains of Arizona and New Mexico greatest densities in warm Pacific Coastal (large form E. d. hellmayri), where E. ham- forests (small form E. d. difficilis) and in mondii is either absent or rare and local. In arid interior highland forests of the Inter- areas of contact between Hammond's Fly- mountain Region and southern Rocky catcher and the similarly sized E. d. diffi- Empidonax difficilis £. d. difficilis E. d. hellmayri Fig. 7. Breeding distributional relationship of Hammond's Flycatcher (principal range outlined) and Western Flycatcher in the contiguous western United States. Relative abundance of three forms of Western Flycatcher is shown by patterns as follows: E. d. difficilis, heavy slanted lines indicate very common to abundant populations, and light slanted lines indicate common to uncommon populations; E. d. hellmayri, black indicates very common to abundant populations, and dotted pattern indicates common to uncommon populations. E. d. insulicola is common to abundant on the California Channel Islands, indicated in black. Scattered localities for F. d. hel- lmayri are shown by circles; note that many of them occur within the main range of E. hammondii. A few local- ities for the latter species, located south of the main range, are omitted. 152 GREAT BASIN NATURALIST MEMOIRS No. 2 cilis in the Pacific Northwest and on the west side of the Cascades, interspecific ter- ritories are defended (Johnson 1966 and un- publ. data). In the Rockies of southern Col- orado the larger E. d. hellmayri evidently overlaps in territories with E. hammondii (Beaver and Baldwin 1975). The important point is that these two species are largely allopatric; considering the bulk of their numbers, this results directly from selection of similar, though not identical, habitats in different major climatic zones. Where they overlap, one species is nearly always com- mon, the other uncommon or rare. Despite the geographic shifts in relative abundance and subtle differences in habitat selection, which presumably are consequences of mil- lenia of competitive interaction, the two species do occur sympatrically in many places. Obviously one species has not limit- ed the geographic range of the other (Fig. 7). Within the genus Empidonax, these two species are separated phylogenetically by a series of other forms of closer relationship. The Hammond's Flycatcher is allied to the Least Flycatcher (E. minimus) and to other species of Boreal derivation (Johnson 1963). The Western Flycatcher, in contrast, is clos- est to Neotropical montane stocks such as E. flavescens and is certainly of southern derivation (N. K. Johnson, unpubl.). Appar- ently such evolutionary backgrounds have permitted extensive convergence in habitat preference and foraging ecology, for in the western United States, E. hammondii and E. difficilis are the commonest Empidonaces of dense coniferous forest. Controls of Boundaries of Species' Ranges The previous accounts offer only the barest introduction to the vast subject of the determinants of borders of breeding dis- tributions of birds. Nonetheless, it is appro- priate to comment briefly on this topic be- cause it lies at the heart of biogeography and because of significant trends in the per- tinent literature. Grinnell (1914, 1917) and Bartholomew (1958), among others, have emphasized manifold controls of distributional bound- aries of animal species, with the importance of particular factors changing along the range perimeter. They favored niche re- quirements and physiologic-climatic prefer- enda as the crucial dictates of species distri- butions; interspecific competition either was not mentioned or its relevance was subordi- nated. In the more recent literature of orni- thological ecology, the limitation of geogra- phic range through competitive exclusion has become a prominent theme (MacArthur 1972, Cody 1974). However, the invocation of competition in this context has been pre- mature. In agreement with Salt (1952), whose classic paper clearly describes how metabo- lism and climate interrelate to control the distributions of three species of finches (Car- podacas), I see little reason to postulate competition as an effective determinant of range borders of species of birds in the In- termountain Region. The question is not whether competition occurs among the many overlapping species here; rather, it is what form does this competition take, and, as such, can it influence range boundaries? From all the evidence I have seen, includ- ing that from the four examples discussed, competition probably has had little in- fluence in the determination of range boundaries of birds. Instead, alterations in density, such as shown for the two species of Empidonax, are the more likely outcomes of competitive interactions. Finally, I sug- gest that it is unwise to conclude that boundaries of tropical species are set by competition (Terborgh and Weske 1975), in the virtual absence of data on the occur- rence of those environmental requisites cru- cially related to the distribution of the bird species in question. Patterns of Avian Speciation in the intermounta1n region In this section I concentrate on what I consider to be active zones of speciation in 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 153 the Great Basin region. These are illustrated by currently interacting populations of closely related taxa, well-differentiated at various levels above that of the incipient subspecies. For those concerned with more subtle examples of racial differentiation in this region, Behle (1963) provides a useful review. Information on speciation zones is of significance to the data previously pre- sented on avifaunal gradients because of the great probability that the same environmen- tal discontinuities in climate and flora ulti- mately are responsible for both. Zones of Intergradation, Disjunction, and Secondary Sympatry Birds demonstrate a particularly wide range of speciation phenomena along the western edge of the Great Basin, in a zone extending south-southeastward from south- ern Oregon to southeastern California and southern Nevada. Indeed, there is evidence from the general distribution and speciation of birds in western North America that the zone just delineated is actually part of a larger biotic separation that extends north- ward to British Columbia and beyond. Here I wish to discuss 16 selected pairs of taxa (Table 4) in the western Intermountain Re- gion that illustrate stages in the speciation process. Figure 8 schematically portrays zones of intergradation and hybridization for nine of these pairs of taxa. Several patterns emerge from study of this wealth of material. The first concerns the striking, although imprecise, geographic coincidence of the several contact zones that link (or divide) coastal and interior populations. Secondly, a large number of species are represented, suggesting that sig- nificant environmental gradients or dis- continuities in the western Great Basin per- vasively influenced the differentiation of birds. Indeed, few other regions in North America can claim a comparable diversity of speciation phenomena within such re- stricted geographic limits. Furthermore, the interacting taxa of nearly every pair occupy different habitats on each side of the contact zone (Table 5). Habitats in the contact areas themselves are typically intermediate between coastal and interior situations. In a major trend illus- trated by six pairs of taxa, the coastal repre- sentative inhabits coniferous forest or oak Table 4. Examples of coastal or Sierra Nevadan and interior taxa that illustrate speciation phenomena along the interface of the Cascades-Sierra Nevada and Great Basin. Gradual primary integradation; racial boundaries weak1: 1. Sphyrapicus thyroideus thyroideus and S. t. na- taliae. Williamson's Sapsucker. Narrow zone(s) of primary or secondary intergradation between representatives of strongly divergent racial complexes: 2. Empidonax difficilis difficilis and E. d. hellmayri2. Western Flycatcher. 3. Eremophila alpestris sierrae and E. a. lampro- chroma. Horned Lark. 4. Aphelocoma coendescens superciliosa and A. c. ne- vadae. Scrub Jay. 5. Psaltriparus minimus californicus and P. m. plumb- eus or P. m. providentialis. Bush-tit. 6. Amphispiza belli canescens and A. b. nevadensis. Sage Sparrow. 7. Passerella iliaca megarhynchus and P. i. fidva or P. i. monoensis. Fox Sparrow. 8. Dendrocopos pubescens turati and D, p. leucurus. Downy Woodpecker. Disjunct allopatry between representatives of strongly divergent racial complexes: 9. Parus inornatus inornatus or P. i. kernensis and P. i. zaleptus. Plain Titmouse. 10. Vireo solitarius cassinii and V. s. plumbeus. Soli- tary Vireo. 11. Vermivora celata lutescens and V. c. orestera. Orange-crowned Warbler. Narrow zone or sympatry and interspecific hybridiza- tion: 12. Sphyrapicus ruber daggetti. Red-breasted Sapsucker and S. varius nuchalis. Red-naped Sapsucker. 13. Junco hyemalis thurberi. Dark-eyed (Oregon) Junco and /. caniceps caniceps. Gray-headed Junco. Disjunct allopatry between species: 14. Pica nuttalli. Yellow-billed Magpie and P. pica hudsonia. Black-billed Magpie. 15. Vermivora ruficapilla ridgwayi. Nashville Warbler and V. virginiae. Virginia's Warbler. 16. Leucosticte tephrocotis dawsoni. Gray-crowned Rosy Finch and L. atrata. Black Rosy Finch. 'Many additional examples of this category could be cited (see Behle 1963). !In each pair of taxa the coastal-Sierran form precedes the interior form. 154 GREAT BASIN NATURALIST MEMOIRS No. 2 3 - OS C —i o -^ ^-i. Dix dale 1914 is nell, Lins •th( c £ "5 r Grin and Swai 5 s c ^ O S o s f S: N Z 13 U « U 0 zu Jai^uS i * ozt,>u I ar 5 J I >' <*- -5 Z c _c o V i&u ^ s c | u X EB S MU -' V 'B U >N 2 J^ a < "5 T3 S — T3 No contact approach a Nev. borde o o -a a a) C o -5 5 8 -5 U as — > z 3 = o o Z 3 2 E E o i s * a- o < "3 j« "3 g -d "B •£ o a " « § '5 | ^ o 1 SiS o 3P ^ s s 5 § | s c B^iX-N C5WJOC/5>OC«&hO a. -s, <1 ."8 II -2 ■* q § o S£ 8 "8 3 ? * 1 -* V. -a ^ o ice 156 GREAT BASIN NATURALIST MEMOIRS No. 2 woodland-chaparral, and the interior race or species lives in pinyon-juniper woodland. In four additional pairs of taxa, the coastal or Sierran form occupies oak-chaparral, co- niferous forest, or brush associated with such forest, and the interior form switches to riparian aspen or willows. Earlier I dis- cussed the latter trend (Johnson 1970). In 5 of the 16 pairs, the coastal and inte- rior representatives are disjunct, with ranges separated by gaps that overlap or straddle many of the zones of intergradation of the 11 pairs of forms that do contact. Evidently the intervening habitat separating the dis- junct forms is unsuitable for these particular species. Seven of the 11 pairs that meet and interact occur less commonly in the contact zone than in the adjacent regions. This clearly suggests that population densities are reduced for several species in and near their zones of intergradation. Thus, the environ- ments at the western border of the Inter- mountain Region are unsuitable for certain species and marginal for others. In any case, the variety of distributional and speciation phenomena represented underscores the steep environmental gradients that influence selection pressures on bird populations in the restricted zone where the Cascade Mountains and Sierra Nevada meet the Great Basin. Munz and Keck (1968), Bald- win (1973), and Miller (1951) describe the significant physiographic, floristic, and cli- matic changes ultimately responsible for the evolution of contrasting avifaunas in this re- gion. Fig. 8. Left, zones of primary intergradation and hybridization (secondary contact), at the western border of the Great Basin, in seven species of birds. Numbers correspond to those in Table 4: 2, Western Flycatcher; 3, Horned Lark; 4, Scrub Jay (two zones); 5, Bush-tit (two zones, one of which is identical with a zone of the Scrub Jay); 6, Sage Sparrow; 7, Fox Sparrow (two zones); and 8, Downy Woodpecker. Bight, localities of sympatry and hybridization between Bed-breasted and Bed-naped Sapsuckers (triangles) and between Dark-eyed (Oregon) and Gray-headed Juncos (dots). 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 157 Physiographic Breaks and lations on the west side of the trough at lo- Avifaunal Boundaries ealities well-removed from their principal centers of distribution from central Nevada A major gap in the mountain systems of eastward. These occurrences provide clear the western United States results from the evidence for the suitability, for at least low trough that begins in southeastern Ore- some interior species, of those montane en- gon, courses southward through western Ne- vironments located directly adjacent to the vada, veers southeastward between Walker southern Cascades and east side of the Lake and the Paradise Range, then contin- Sierra Nevada. But these same ranges also ues through the Death Valley area and host many Cascadian and Sierran forms that beyond into the Mojave and Colorado have colonized from the west. Thus, pecu- deserts. With such a geographic position, it Harly intermixed avifaunas breed in all four would be easy to assume that this trough mountain systems (Table 6). One of the four separates the Boreal avifaunas of the Cas- ranges, the Sweetwaters, has only a minor cades and Sierra Nevada from those of the representation of interior forms; the others interior. However, this is not the case, for a have numerous Great Basin-Rocky Mountain host of interior races and species of birds species. All of the other mountain systems occur westward to the Warner, Wassuk, along the Nevada-California border, the Sweetwater, and White mountains. These Carson Range, Pine Nut Mountains, Pine interior forms therefore maintain popu- Grove Hills, and Glass Mountain, lack inte- Table 6. Intermixture of Cascade-Sierra Nevadan and interior components1 in Boreal avifaunas of mountain ranges2 along the western edge of the Great Basin. Cascade Mountain-Sierra Nevadan Warner Wassuk Sweetwater White Forms Mountains Range Mountains Mountains Dendragapus obscurus sierrae X - X X Sphyrapicus ruber daggetti X X X - Cyanocitta stelleri frontalis X X X X Vireo soliturius cassinii X - - - Vermivora ruficapilla ridgwayi X - - - Wilsonia pusilla chryseola - - X - Leucosticte tephrocotis dawsoni - - X X Junco hyemalis thurberi X X X X Passerella iliaca monoensis2 - X X - Great Basin-Rocky Mountain Forms Selasphorus platycercus - X - X Sphyrupicus varius nuchalis X X X X Dendrocopos pubescens leucurus X - - - Empidonax difficilis hellmayri X - - -* Vireo soliturius plumbeus - X X X Vermivora celata orestera X - - X Vermivora virginiae - X - X Wilsonia pusilla pileolata X - - X Junco caniceps cuniceps - - - X Passerella iliaca ftdva X — — — Passerella iliaca canescens - - - X 'Emphasis is on forms of contrasting Sierra Nevadan versus interior distribution. Mam undifferentiated species of widespread distribution in the west- ern United States are not considered here. 'Sources of data: Warner Mountains, Miller (1951) and Johnson (1970); Wassuk Range, original data of author and few records from Linsdale (1936); Sweetwater Mountains, specimens in MVZ and original data of author; White Mountains, Grinnell and Miller (1944) and Miller and Russell (1956). The avifaunal relationships of this form are uncertain. 'Miller and Russell (1956) report E. d. difficilis in late June, but the bird was not certainly nesting even though the specimen was in breeding condi- tion. 158 GREAT BASIN NATURALIST MEMOIRS No. 2 rior forms; their Boreal avifaunas are entire- ly comprised of Sierra Nevadan and wide- spread species. Because the most distinctive interior species and racial complexes, such as Selasphorus platycercus, Vermivora vir- giniae, and the ivoodhonseii Group of Aph- elocoma coerulescens, extend all the way from eastern California to Colorado, there is no Boreal avifauna typical only of the Great Basin section of the interior. The aforementioned distributional evi- dence thus supports the principal conclusion derived from the location of contact zones, namely, that the most trenchant separation of coastal-Sierran versus interior stocks oc- curs in a zone of varying width that runs southward just east of the crest of the Cas- cades then southeastward along the inter- face of the Sierra Nevada and Great Basin. Presumably it is the beginning of the pre- cipitation shadow and the consequent shift from a coastal to an interior continental cli- mate that provide the profound environ- mental transition responsible for this zone. The low desert trough to the east does not represent a significant avifaunal barrier. Epilogue.— Avian geography in the Inter- mountain Region continues to progress, al- beit cautiously, from the descriptive to the analytic phase. With only crude patterns now discernible, much remains to be learned and interpreted. But, heeding E. O. Wilson's (1970) reminder that ". . .biogeo- graphy is far and away the most difficult of all the biological sciences," I am inclined to treat with great kindness even those tenta- tive patterns that so far have surfaced from the chaos. Acknowledgments Much of the field work on which the present synthesis is based was supported by the Committee on Research, University of California, Berkeley. Mercedes S. Foster read a draft of the manuscript and offered valuable comments. Gene M. Christman drafted the final versions of the illustrations. I am indebted to all these individuals for their kind assistance. Literature Cited American Ornithologists' Union. 1957. Checklist of North American birds, fifth ed. Amer. Orni- thol. Union, Baltimore, Maryland. 1973. Thirty-second supplement to the American Ornithologists' Union checklist of North American birds. Auk 90: 411-419. Baldwin, J. L. 1973. Climates of the United States. U.S. Dept. Commerce, Government Printing Office, Washington, D.C. Bartholomew, G. A., Jr. 1958. The role of phys- iology in the distribution of terrestrial verte- brates, p. 81-95. In: C. L. Hubbs (ed.), Zoogeo- graphy. Amer. Assoc. Adv. Sci., Washington, DC. Beaver, D. L., and P. H. Baldwin. 1975. Ecological overlap and the problem of competition and sympatry in the western and Hammond's flycatchers. Condor 77: 1-13. Behle, W. H. 1942. Distribution and variation of the horned larks (Otocoris alpestris) of western North America. Univ. Calif. Publ. Zool. 46: 205-316. 1963. Avifaunistic analysis of the Great Basin region of North America. Proc. Inter. Ornith. Congr. 13: 1168-1181. Behle, W. H., and M. L. Perry. 1975. Utah birds: checklist, seasonal and ecological occurrence charts, and guides to bird finding. Utah Mus. Natural Hist., Univ. Utah, Salt Lake City. Burleigh, T. D. 1972. Birds of Idaho. The Caxton Printers, Ltd., Caldwell, Idaho. Cody, M. L. 1974. Competition and the structure of bird communities. Princeton Univ. Press, Princeton, N.J. Critchfield, W. B., and G. L. Allen- baugh. 1969. The distribution of Pinaceae in and near northern Nevada. Madrono 19: 12-26. Grinnell, J. 1914. Barriers to distribution as re- gards birds and mammals. Amer. Naturalist 48: 248-254. 1917. Field tests of theories concerning dis- tributional control. Amer. Naturalist 51: 115-128. Grinnell, J., J. Dixon, and J. M. Linsdale. 1930. Vertebrate natural history of a section of northern California through the Lassen Peak region. Univ. Calif. Press, Berkeley. Grinnell, J., and A. H. Miller. 1944. The distri- bution of the birds of California. Pacific Coast Avif. 27: 1-608. Hayward C. L., M. L. Killpack, and G. L. Bichards. 1963. Birds of the Nevada Test Site. Brigham Young Univ. Sci. Bull., Biol. Ser. 3(1): 1-27. Howell, T. B. 1952. Natural history and differen- tiation in the yellow-bellied sapsucker. Condor 54: 237-282. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 159 Johnson, N. K. 1956. Birds of the pinon association of the Kawich Mountains, Nevada. Great Basin Nat. 16: 32-33. 1963. Biosystematics of sibling species of fly- catchers in the Empidonax haminondii-oberhol- seri-wrightii complex. Univ. Calif. Publ. Zool. 66: 79-238. 1965. The breeding avifaunas of the Sheep and Spring ranges in southern Nevada. Condor 67: 93-124. 1966. Morphologic stability versus adaptive variation in the Hammond's flycatcher. Auk 83: 179-200. 1970. The affinities of the Boreal avifauna of the Warner Mountains, California. Occas. Pap. Biol. Soc. Nevada 22: 1-11. 1973. The distribution of Boreal avifaunas in southeastern Nevada. Occas. Pap. Biol. Soc. Ne- vada 36: 1-14. 1974. Montane avifaunas of southern Ne- vada: historical change in species composition. Condor 76: 334-337. 1975. Controls of number of bird species on montane islands in the Great Basin. Evolution 29: 545-567. 1976. Breeding distribution of Nashville and Virginia's warblers. Auk 93: 219-230. Johnson, N. K., and K. L. Garrett. 1974. Interior bird species expand breeding ranges into south- ern California. W. Birds 5: 45-56. Johnson, B. E. 1975. New breeding localities for Leucosticte in the contiguous western United States. Auk 92: 586-589. Linsdale, J. M. 1936. The birds of Nevada. Pacific Coast Avif. 23: 1-145. 1938. Environmental responses of vertebrates in the Great Basin. Amer. Midland Naturalist 19: 1-206. MacArthur, B. H. 1972. Geographical ecology. Harper and Bow, New York. Miller, A. H. 1941. Speciation in the avian genus Junco. Univ. Calif. Publ. Zool. 44: 173-434. 1946. Vertebrate inhabitants of the pinon as- sociation in the Death Valley region. Ecology 27: .54-60. 1951. An analysis of the distribution of the birds of California. Univ. Calif. Publ. Zool. 50: 531-644. Miller, A. H., and W. C. Bussell. 1956. Distributional data on the birds of the White Mountains of California and Nevada. Condor 58: 75-77. Munz, P. A., and D. D. Keck. 1968. A California flora. Univ. Calif. Press, Berkeley. Orr, B. T., and J. Moffitt. 1971. Birds of the Lake Tahoe region. California Academy of Sciences, San Francisco, Calif. Pitelka, F. A. 1951. Speciation and ecologic distri- bution in American jays of the genus Aphelo- coma. Univ. Calif. Publ. Zool. 50: 195-464. Salt, G. W. 1952. The relation of metabolism to climate and distribution in three finches of the genus Carpodacus. Ecol. Monogr. 22: 121-152. St. Andre, G., H. A. Mooney, and B. D. Wright. 1965. The pinyon woodland zone in the White Mountains of California. Amer. Midi. Naturalist 73: 225-239. Swarth, H. S. 1914. The California forms of the genus Psaltriparus. Auk 31: 499-526. Taylor, W. P. 1912. Field notes on amphibians, reptiles, and birds of northern Humboldt Coun- ty, Nevada. Univ. Calif. Publ. Zool. 7: 319-436. Terborgh, J., and J. S. Weske. 1975. The role of competition in the distribution of Andean birds. Ecology 56: 562-576. Wilson, E. O. 1970. Facts of zoogeography. Science 168: 1193-1194. EXPLOSIVE EVOLUTION OF PERENNIAL ATRIPLEX IN WESTERN AMERICA1 Howard C. Stutz2 .Abstract.— New habitats opened up in western North America since the recession of Lake Bonneville and Lake Lahontan have admitted a host of new species of Atriplex. Every known evolutionary force is operating and at accelerated paces. Autoploidy appears to be more common than in any other reported group of plants. Natu- ral hyoridization between closely related species has provided a wealth of fertile segregants from which new adaptive types have been and are being selected. Hybrids between more distantly related species are sterile and some appear to have given rise to fertile alloploid derivatives. All evidence points to a center of origin for Atriplex in northern Mexico. The numerous species which have migrated northward into western United States and Canada were apparently able to do so because attributes ac- quired to make them adaptive in the hot dry deserts of Mexico were characteristics which, uniquely, also pre- adapted them for colder climates and alkaline clay soils to the north. Woody species such as Atriplex canescens and A. con ferti folia hybridize rather easily with herbaceous pe- rennial species such as A. cuneata, A. gardneri, A. corrugata, and A. obovata. Since most such hybrids are at least partly fertile and produce F2 segregants of both woody and nonwoody types, the genetic basis for the accumulation of wood is apparently rather simple. According to Antevs (1955), Rroecker and Kaufmann (1965), Morrison (1965), Russell (1885), and others, much of the land surface of western Utah and northwestern Nevada was covered with ice and water as recently as 10,000-12,000 years ago. During the sub- sequent rapid disappearance of Lake Bonne- ville and Lake Lahontan and the attending desertizing of surrounding valleys, numerous salt playas and alkali flats were formed. Very few plants have been capable of sur- viving the severe physiological drought which characterizes such habitats. Com- pounded by attending severe climatological drought, which in many places is nearly ab- solute, the number of adaptive plants and animals is even further minimized. Indeed, in many areas, islands of dry, saline, mud hills and alkali playas are still completely uninhabited. Such areas represent some of the few last frontiers on earth yet to be ex- ploited by living organisms. The principal pioneers which reach out farthest into these sterile desert islands are plants belonging to the family Chenopo- diaceae. The entire family appears to pos- sess, uniquely, characteristics which permit the accommodation of this double challenge of physiological and climatological drought. At the borders of every sterile, empty island in these saline deserts some member of this remarkable family is at the last frontier. In bottomlands it is usually Sarcobatus or Saaeda or Salicornia or Allenrolfea. On dryer sites it is Graijia or Ceratoides or Atriplex or Sa Isola. Most chenopod genera are highly special- ized with very few adaptive variables. Only two species of Sarcobatus have been de- scribed: S. vermiculatus (Hook.) Torr and S. 'This research was supported by USDA Forest Service, Intermountain Forest and Range Experiment Station, Agreement 12-11-204-31 No. 8. 'Department of Botany and Range Science, Brigham Young University, Provo, Utah 84602. 161 162 GREAT BASIN NATURALIST MEMOIRS No. 2 baileyi Cov. There are only two species of Grayia: G. spinosa (Hook.) Moq. and G. brandegei Gray. There is only one species of Ceratoides (C. lanata (Pursh) J. T. How- ell) in North America and one of Cycloloma (C. atriplicifolium (Spreng.) Coult.). But, in contrast, Atriplex is awesomely genetically rich, consisting of numerous species and va- rieties, many of which are still unnamed. Some Atriplex species are very narrowly endemic, others are phenomenally wide- spread. Atriplex navajoensis Hanson is con- fined to a restricted area near Navajo Bridge in northern Arizona, whereas Atri- plex canescens (Pursh) Nutt. extends all the way from Montana southward, deep into Mexico. Some species such as Atriplex gar- rettii Rydb. are remarkably uniform throughout their entire distribution; others, such as Atriplex confertifolia (Torr. and Frem.) S. Wats, are highly flexible with nu- merous ecotypes, chromosome races, and subspecies. So genetically rich is Atriplex in western America and so new and variable are the environs which it occupies that we are per- mitted to witness the evolutionary process at an unprecedented rate. Every known strategy for speciation is evident, some of which, such as autoploidy, have never been reported to be nearly so significant in the evolution of other plant groups. Interspecific hybridization followed by in- trogression, segregation, or alloploidy is common; mutations, drift, and autoploidy are found in nearly every species. Inter- pretations are easy in some groups, but in others the evidences are rather subtle and difficult to interpret. Evolutionary Hot Spots Atriplex is abundant throughout all of western America. One or more of the evo- lutionary forces can be witnessed in nearly every population, but there are a few par- ticular sites where speciation is especially rapid. One of the richest evolutionary sites lies in northeastern Mexico in a radius of about 200 miles around Monterey. Here there are numerous endemic diploid and tetraploid species, and numerous hybrid derivatives. As pointed out by Johnston (1941), Atriplex stewartii I. M. Johnston is an apparent de- rivative of the hybrid Atriplex canescens x A. acanthocarpa (Torr.) S. Wats. It has thin, broad leaves with undulating margins, much like the leaves of A. acanthocarpa, and four-winged fruits which, although dimin- utive, are much like those of A. canescens. West of San Roberto is another segregant from this hybrid but, in this case, with A. canescens-type leaves and A. acanthocarpa- type fruits. Even another derivative from this hybrid grows in very heavy gypsum soils southwest of Cuatro Cienegas. It is much like A. acanthocarpa in many respects of habit and leaf but has a strong tendency for wings on an otherwise burrlike fruit. Atriplex prosopidium I. M. Johnston is a handsome, tall, four-winged shrub with a deep blue cast which probably came from introgression of A. obovata Moq. into A. ca- nescens. It is narrowly endemic to a small area around Monclova and Ocampo. A much more widespread putative derivative of this same origin covers thousands of acres south and west of Cuatro Cienegas. It is a short-statured, heavy bush, much utilized by sheep and cattle in that area. A second geographic center in which hybridization has played, and is now play- ing, a major role in the origin of new spe- cies of Atriplex is in north central Nevada. Valleys in this region run north and south between series of mountain ranges. Each valley is approximately 30 to 60 miles across and 100 to 400 miles long and usual- ly opens up to common basins to the north and to the south. This peculiar phys- iography permits species to migrate from common populations into isolated valleys where independent speciation can occur and unique species can emerge. And they have. In almost every valley unique gen- otypes have become established. Further- more, fortuitous contacts between north and south immigrants as well as fortuitous in- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 163 troductions of particular combinations of species from either the north or the south has provided many of the valleys with new rich opportunities for interspecific hybridi- zation followed by segregation, in- trogression, and polyploidy. Atriplex canescens has been particularly contributive in this setting. In western Utah and northern Nevada it has hybridized with A. tridentata Kuntze yielding a host of new variables. This is particularly unusual be- cause A. canescens in this area is tetraploid whereas A. tridentata is hexaploid. However numerous viable seeds are produced by the hybrid in nature, providing a host of segre- gating progeny. Some of these have recov- ered high fertility and chromosomal regu- larity at the tetraploid level but are phenotypically, and therefore genetically, strongly modified by A. tridentata. Con- versely, some segregants have settled down chromosomally at 2n = 54 and are phenotypi- cally like A. tridentata, introgressed with A. canescens. Still others are not yet stabilized, either genetically or chromosomally. Each of these permutations have also been ob- tained in the experimental nursery at Brig- ham Young University from seeds collected from natural hybrids. In and around Battle Mountain, Nevada, thousands of acres are occupied by a highly variable population of Atriplex which ap- pears to have been derived from this par- entage. The plants are strongly A. tridentata in many features but tendencies for, winged fruits, earlier maturation, and woodiness all suggest introgression from A. canescens. This population appears to have become chromosomally stabilized at 27 pairs but is still unstabilized genetically. In many of the valleys of west central Nevada, Atriplex canescens appears to pos- sess genetic attributes of A. falcata (M.E. Jones) Standi. Since these A. canescens pop- ulations are hexaploid, whereas most others in western America are tetraploid (Stutz et al. 1975), and, since A. falcata has so far proven to be always diploid, these hex- aploid A. canescens populations are prob- ably allohexaploids derived from the hybrid of tetraploid A. canescens x diploid A. fal- cata. The principal difficulty with this in- terpretation, however, is the abundant vari- ation within these hexaploid populations. A monophyletic hexaploid origin should have provided a very uniform product. Con- sequently, if the parentage has been cor- rectly interpreted, then a polyphyletic ori- gin is required, either initially or from among partially fertile segregants of the original hybrid or else abundant continuing hybridization between the new allohexa- ploid and one or both of the parents, fol- lowed by subsequent introgression. Another highly variable hexaploid popu- lation occupies extensive valleys south of Eureka, Nevada. These are much less like A. canescens than are those described above which occupy valleys to the west. If these, too, are allohexaploid derivatives from the hybridization of tetraploid A. canescens and diploid falcata, considerable segregation must have occurred prior to the event of chromosome doubling because these plants show a minor influence of A. canescens and a major A. falcata input. Certainty about whether or not these are the correct interpretations must await fur- ther study and experimental breeding. But whatever the ancestry, the fact remains that these partially isolated valleys in central Nevada are hot beds of Atriplex evolution. Nearly every valley possesses unique gen- otypes. In many valleys the variation is still rampant and has not yet settled down to a predominant adaptive theme. It is an evolu- tionary arena in which speciation is pro- ceeding at unprecedented rates. Eastern Utah is another region in which both polyploidy and interspecific hybridiza- tion have furnished many new species and varieties. The most common interspecific- hybrids are those derived from A. cuneata A. Nels x A. canescens and A. cuneata x A. confertifolia. The latter is highly sterile but a few segregants are to be found in nature, and I have now obtained several seedlings from many thousands of hybrid "seeds." 164 GREAT BASIN NATURALIST MEMOIRS No. 2 The hybrid A. canescens x cuneata, how- ever, is, in some places, quite fertile and produces extensive segregating populations. Immediately west of Hanksville, Emery Co., Utah, about thirty acres is dominated by hundreds of segregants from this parentage. Since A. cuneata is usually a suffrutescent shrub, dying back almost entirely each year to the ground, whereas A. canescens is a woody shrub, accumulating wood year after year, it was interesting to find both growth forms represented among the segregants, suggesting a rather simple genetic basis for this difference. Numerous other natural interspecific- hybrids have been found in eastern Utah, e.g., A. confertifolia x A. garrettii, A. confer- tifolia x A. corrugata S. Wats., and A. con- fertifolia x A. canescens, but extensive suc- cessful segregants have not yet been noted. Probably the most significant products of interspecific hybridization in all Atriplex is that produced from hybridization of A. ca- nescens with A. gardneri (Moq.) Dietr. In northern Wyoming and southern Montana these two species are often intimately sympatric. Hybridization between them is common, attended by a rich array of segre- gant products. From such hybrid swarms have come several highly adaptive com- binations, one of which appears to be the widespread, herbaceous form of A. canescens which grows in the heavy clay soils on the banks of the Missouri River and its tributaries throughout most of Montana, southern Alberta, northern Wyoming and North and South Dakota. It was this form which was collected in 1804 by Lewis and Clark at Big Bend near Chamberlain, South Dakota, and which is the type for A. canes- cens. Because of the many differences which distinguish it from the typical taller, more woody, well-known four-wing saltbush of the Intermountain West and northern Mexico, a nomenclatural revision will be re- quired. Origins It is difficult to be certain of the center of origin of species in a genus as genetically rich as Atriplex. According to Vavilov (1926), it would be in one of the areas in which there is abundant variation. How- ever, in Atriplex this may not be a good in- dex for origins. As pointed out above, most of the evolutionary hot-spots are of very re- cent vintage and possess abundant variation not because of antiquity but rather because of recent availability of new explorable hab- itats. Also, unacceptable are theories such as those of Matthew (1939) which suggest that centers of origins can be detected by the presence of advanced, progressive species. This, too, would place the center of origin in northern Utah and Nevada or eastern Utah, where evolution is currently explosive and where, but a few thousand years ago, much of the area was covered with water. In view of the widespread severe ecologi- cal changes during Pleistocene and Holo- cene in all of western America north of Mexico, in view of the numerous affinities between perennial Atriplex species of north- ern Mexico and those to the north, and in view of the numerous species in southern states and Mexico such as A. acanthocarpa (Torr.) S. Wats., A. obovata Moq. A. pohj- carpa (Torr.) S. Wats., A. hipnenelytra (Torr.) S. Wats., and A. torreyi (S. Wats.) S. Wats, which have no homologues nor even analogues in northern states or Canada and also because there is a paucity of perennial Atriplex species south of San Luis Potosi, northeastern Mexico appears to be the most likely area from which most contemporary perennial, North American species of Atri- plex arose. Adaptation to the hot dry cli- mate and gypsiferous soils of this area con- ceivably could have preadapted many forms for occupation of the xeric saline areas which became available to the north during late Pleistocene. It may have involved bio- logical adaptation as simple as the capacity for accumulating and tolerating salts at an unusually high concentration. Ostensibly the attending high osmotic pressure of such an adaptation could function as an "antifreeze" in colder northern climates as well as a 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 165 mechanism for accommodating drought, both physiological and climatological. No other common denominator is yet apparent. Such an origin is also fully supported by other lines of evidence. Leaves of all pe- rennial Atriplex species, so far examined, have the "Kranz" type leaf anatomy which appears to be always associated with the C4 photosynthetic pathway (Downton 1971, Polya and Osmond 1972). Even Atriplex fal- cata which now is confined solely to north- ern latitudes and other diploid species as far north as Edmonton, Alberta, still display Kranz type anatomy. If, as is generally maintained, C4 photosynthesis is an adapta- tion to hot, bright desert conditions, (Down- ton 1971, Mooney et al. 1974, Smith 1976, and others) it is difficult to accept that these northern species arose in situ under existing climatic conditions. However, a southern hot-desert origin for a species now completely restricted to northern climates is also difficult to accept. But if the Kranz anatomy is as reliable an index to C4 photo- synthesis as has been claimed, and if C4 photosynthesis is as restricted to hot climate plants as has been suggested, then there are really no other available explanations. And if we accept all of that then we are forced to accept, in either case, migration and evo- lutionary tempos of very unusual rates. The only apparent northerly migration lanes were right through Utah and Nevada, and these were mostly under water only 10 or 12 thousand years ago. If this migration were earlier, it must have been during the interstadial between the Bull Lake and Pinedale pluvials, and the current distribu- tion of northern species as far south as northern Nevada and Utah would be a re- cent subsequent southerly migration. Since most of Alberta and Saskatchewan was completely covered with ice and water as recently as 20,000 years ago (Bayrock 1969, Christiansen 1971), Atriplex species which are now restricted to areas in these provinces as far north as Edmonton and Peace Biver must have immigrated from the south since that time. Also, since consid- erable time would be required for the wasting and disappearance of the ice sheets, desert conditions conducive to Atriplex mi- gration would not have been available until even much later. And, since most of Utah and Nevada were still mostly covered with ice and water as recently as 10,000 years ago (Bissell 1963, Morrison 1965), there are not very many time periods left in which the required major desert could have devel- oped. In fact, it appears that there are only two possibilities: (1) about 10,000 to 12,000 years ago, just prior to the last major Bon- neville pluvial or (2) about 5,000 years ago during the Altithermal period of post Bon- neville. In view of the great distance which migrating species would need to traverse to get from southern deserts to far northern latitudes, plus time for radiation into the numerous existing adaptive ecotypes and in situ synthesis of polyploid derivatives, nei- ther of these time intervals are particularly appealing. But since there appears to be no available alternative it must have happened during one or perhaps both of these periods. The tempo of migration and evolution would have necessarily been very rapid. But since it is very rapid today, perhaps it is not so unrealistic after all. According to Morrison (1965), a distinct period of dessication separated the Bonne- ville and Draper pluvials about 12,000 years ago. It was a relatively brief period during which the Grantsville soil was deposited at levels as low as 4250 feet. This is just 50 feet above the current level of Great Salt Lake so would indicate desert conditions perhaps as severe as those existing today. Also Eardley (1962) found a huge bed of Glauber's salt (Na2SO4«10H2O) from 15 to 25 feet below the present-day bottom of Great Salt Lake, 9lA miles across and 32 feet thick, the surface of which dated at 11,600 ± 400 years B.P. Further evidence of a major drought 10,000-12,000 years ago is furnished by cores taken from Searles Lake which show an 18-feet- thick layer of salt of that age (Smith 1962). Increasing frequencies of Atriplex confer- 166 GREAT BASIN NATURALIST MEMOIRS No. 2 tifolia in wood-rat middens in the Mojave Desert beginning about 17,000 years ago (Wells 1976) may also be counted as evi- dence for extant desert conditions in west- ern North America prior to the last Bonne- ville pluvial. Wells also furnished evidence which suggests that the Chihuahuan Desert may have originated less than 11,500 years ago. Widespread desert conditions may have lasted for only a few hundred or perhaps a few thousand years but were probably suffi- ciently extensive to permit Atriplex and oth- er desert plants to migrate from the south. Some of the Atriplex species which are now endemic to the northern latitudes are, uniquely, successfully competitive with grasses and forbs with which they grow. Elsewhere Atriplex is conspicuously a poor competitor and is successful on harsh sites primarily because nothing else can grow there. Consequently the northern forms have apparently become secondarily adapt- ed for competitiveness in sites which are considerably more mesic than those occu- pied by other species of Atriplex further to the south. Since such adaptation is likely genetically somewhat complex, it probably required considerable time to arise. This is further argument that migration from the south must have occurred at an early period rather than during the Altithermal of only 5,000 to 6,000 years ago. The pluvial period following the extant drought 10,000 to 12,000 years ago appar- ently wiped out all northern forms of Atri- plex except those which were adaptively competitive with other mesophytes. At least one of these northern Atriplex species, near Edmonton, Alberta, is diploid so probably represents a product of primary evolution involving principally genetic drift and the acquisition of new adaptive mutations. Considerable evidence is now accumulat- ing which suggests that desert conditions were also widespread during the Altither- mal period. Scott (1965) reports mollusks dated at 5500 B.P., buried 40 feet under eo- lin sand near Denver, Colorado, and Smith (1962) reports 40 feet of salt of altithermal age deposited in Lake Searles in eastern California. Whether the warm Altithermal period developed synchronously throughout all of western America to provide one huge desert or whether deserts developed in dif- ferent areas at different times is not known but neither is it critical to Atriplex evolu- tion and migration. A single widespread desert would, of course, be simpler to com- prehend but stepwise movement from one area to another could have been just as ef- fective. The principle conclusion in either case is the same. Atriplex must have mi- grated from the south, over thousands of miles, across terrain which has subsequently become so modified that intervening forms have become completely extinct. Ancestral types, if they still exist, are so very different from those to the north which have evolved from them that evidences of such ancestry are not at all apparent. Although the climate of northeastern Mexico has apparently changed very little during the past 30,000 to 40,000 years (Meyer 1973) in comparison to the tu- multuous surface changes to the north, Atri- plex growing there would have certainly continued to evolve. However, if the north- ern derivatives migrated from the south during the recent Altithermal period, many original ancestral types would predictably still be present. But none are to be found, a fact which suggests that migration must have been much earlier. Wells (1966) has reported Atriplex canes- cens fossils at Burro Mesa in the Chihua- huan Desert which date at 36,000 years B.P. However, today only polyploid forms of this species have been found in Mexico (Stutz et al. 1975). The only population of diploid Atriplex canescens yet found is on sand dunes in central Utah and may repre- sent the only relic stand of the ancestral form. However it is so genetically dis- tinctive and so restricted in its distribution, it is almost certainly a derived form, consid- erably altered from its progenitors. Many other diploid species of Atriplex scattered throughout the Intermountain 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 167 West are likewise geographically disjunct from any apparent Mexican relative. Most of them have narrow restricted distributions so were probably left behind as adaptations to unique situations during and after the migration rampage. All of this argues strongly for greater an- tiquity in Atriplex migration than can be af- forded by the Altithermal period and con- sequently supports other evidences for widespread desert conditions prior to the last Bonneville pluvial, of a magnitude suf- ficient to accommodate migration and evo- lution every bit as rapid as that which we witness today in the wake of new ecological sites exposed by post-Bonneville climatic changes. As new ecological permutations now emerge, encroaching evolutionary fronts ex- ploit them. Hence the entire Intermountain West today is once again an evolutionary arena in which new species are emerging and migrating at phenomenal rates. Moun- tains which are islands for mesophytes are isolation barriers for xerophytes, and the valleys which are isolation barriers for mesophytes are islands for the xerophytes. As these isolation barriers are surmounted by chance migrants, or adaptive deviants, new waves of evolutionary turbulence evoke yet other genetic permutations which accommodate even other challenges. It would be difficult to imagine a more dy- namic arena wherein to witness the evolu- tionary process and it would also be diffi- cult to find a genus better fitted for adaptively responding to it than Atriplex. Literature Cited Antevs, E. 1955. Geologic climatic dating in the west. Amer. Antiquity 20: 317-335. Bayrock, L. A. 1969. Incomplete continental glacial record of Alberta, pp. 99-103. In: H. E. Wright, Jr. (ed.), Quaternary geology and climate. Na- tional Academy of Sciences, Washington, D.C. Bissell, H. J. 1963. Lake Bonneville: geology of southern Utah Valley, Utah. U.S. Geol. Surv. Prof. Pap. 257-B: 101-130. Broecker, W. S., and A. Kaufman. 1965. Radiocarbon chronology of Lake Lahontan and Lake Bonneville II, Great Basin. Bull. Geol. Soc. Amer. 76: 537-566. Christiansen, E. A. 1971. Tills in southern Sas- katchewan, Canada, pp. 167-183. In: R. P. Goldthwait (ed.), Till, a symposium. Ohio State University Press, Columbus. Downton, W. J. S. 1971. Adaptive and evolutionary aspects of C4 photosynthesis, pp. 3-17. In: M. D. Hatch, C. B. Osmond, and R. A. Slayter (eds.), Photosynthesis and respiration. Wiley, New York. Eardley, A. J. 1962. Glauber's salt bed west of Promontory Point, Great Salt Lake, Utah. Geol. Mineral. Surv. Spec. Stud. 1: 1-12. Johnston, I. M. 1941. New phanerogams from Mex- ico IV. J. Arnold Arbor. 22: 110-124. Matthew, W. D. 1939. Climate and evolution. 2d ed. New York Academy of Sciences, New York. Meyer, E. R. 1973. Late Quaternary paleoecology of the Cuatro Cienegas Basin, Coahuila, Mexi- co. Ecology 54: 982-995. Mooney, H. A., J. H. Troughton, and J. A. Berry. 1974. Arid climates and photo- synthetic systems. Carnegie Inst. Wash. Year Book 73: 793-805. Morrison, R. B. 1965. Quaternary geology of the Great Basin, pp. 265-285. In: H. E. Wright, Jr., and D. G. Frey (eds.), The Quaternary of the United States. Princeton University Press, Princeton, New Jersey. Polya, G. M., and C. D. Osmond. 1972. Photophosphorylation by mesophyll and bundle sheath chloroplasts of C4 plants. PI. Physiol. (Lancaster) 49: 267-269. Russell, I. C. 1885. Geological history of Lake La- hontan. A Quaternary lake of northwestern Ne- vada. U.S. Geol. Surv. Monogr. 11: 1-288. Scott, G. R. 1965. Nonglacial Quaternary geology of the southern and middle Rocky Mountains, pp. 243-254. In: H. E. Wright, Jr., and D. G. Frey (eds.), The Quaternary of the United States. Princeton University Press, Princeton, New Jersey. Smith, B. N. 1976. Evolution of C4 photosynthesis in response to changes in carbon and oxygen concentrations in the atmosphere through time. Biosystems 8: 24-32. Smith, G. I. 1962. Subsurface stratigraphy of late Quaternary deposits, Searles Lake, California, a summary. U.S. Geol. Surv. Prof. Pap. 450-C: 65- 69. Stutz, H. C, J. M. Melhy, and G. K. Livingston. 1975. Evolutionary studies of Atriplex: a relic gigas diploid population of Atriplex canescens. Amer. J. Bot. 62: 236-245. 168 GREAT BASIN NATURALIST MEMOIRS No. 2 Vavilov, N. I. 1926. Studies on the origin of culti- huahuan Desert. Science 153: 970-975. vated plants. Trudy Prikl. Bot. Selekc. 16(2): 1- 1976. Macrofossil analysis of Wood Rat (Neo- 248. toma) middens as a key to the Quaternary vege- Wells, P. V. 1966. Late Pleistocene vegetation and tational history of arid America. Quaternary degree of pluvial climatic change in the Chi- Res. 6: 223-248. DISTRIBUTION AND PHYLOGENY OF ERIOGONOIDEAE (POLYGONACEAE) James L. Reveal1 Abstract.— Eriogonoideae is a subfamily of the knotweed family, Polygonaceae, endemic to the New World, and is composed of 14 genera and perhaps 320 species. It differs primarily from the other members of Polyg- onaceae in lacking well-defined sheathing stipules or ochrea. The species of Eriogonoideae vary from tiny, fragile annuals to herbaceous perennials, low subshrubs or shrubs to large and often arborescent shrubs. The seemingly most primitive extant genus of the subfamily is Eriogonum (247 species), which is widespread in central North America. A series of genera are closely related to Eriogonum, and probably have evolved directly from Eriogo- num. These genera are Oxytheca (9 species) of the western United States, and Chile and Argentina of South America; Dedeckera and Gilmania, both monotypic genera of the Death Valley region of California; Stenogonum (2 species) of the Colorado Plateau and adjacent areas of the Rocky Mountain West; Goodmania and Hollisteria, 2 monotypic genera of central and southern California; and Nemacaulis, a monotypic genus of the southwestern United States and northwestern Mexico. A second major complex of genera also probably evolved from Eriogo- num. In this group, the most elementary genus is Chorizanthe (about 50 species), in which the extant perennial members of the genus are perhaps evolutionarily the oldest taxa of the subfamily. These perennials are restricted to Chile, while in the western United States and northwestern Mexico of North America, only annual species are found. Mucronea (2 species) of California and Centrostegia (4 species) of the southwestern United States and northwestern Mexico are clearly related to Chorizanthe. In a somewhat intermediate position between the Eriogo- num complex and the Chorizanthe complex— but still more closely related to the latter than the former— is the genus Lastarriaea (2 species) found in California, Baja California, and Chile. All of these genera belong to the tribe Eriogoneae. A second tribe, Pterostegeae, contains only 2 discordant, monotypic genera: the shrubby per- ennial genus Harfordia of Baja California and the more widespread annual, Pterostegia, of the western United States. While time and evolution have obscured the relationships between Eriogoneae and Pterostegeae, the affi- liations among the various genera of the tribes can be ascertained to some degree. The geographical center of origin of the subfamily may have been in a subtropical climate, with the differentiation of modern-day genera oc- curring in temperate, xeric regions of North America. The origin of Chorizanthe was an ancient development, with the migration of the primitive perennial members into South America in the Tertiary. The subsequent de- velopment of the annual habit, and migration of annual species of Eriogonoideae into South America has prob- ably occurred in the Quaternary. The intermediate stages of evolutionary development of the genera and species of the subfamily occurred in a habitat similar to the pinyon-juniper woodlands of the Great Basin, while evolu- tion of the more advanced genera and species has occurred in xeric grasslands, chapparral scrub, or xerophytic "hot desert" communities. Introduction In considering the action of evolutionary processes . . . mainly in the Northern Hemisphere. It con- both extinction and extensive alterations of geographic tams m important agricultural and horti- and ecological distribution patterns must be recognized. , , .-it.- n (Stebbins 1974: 37). cultural species in addition to many well- known and troublesome weeds. The vast Polygonaceae Juss. is a large, temperate majority of the plants are small, inconspic- or subtropical family of flowering plants uous members of the world's vascular plant found throughout much of the world, but flora, and they can claim few positive attri- 'Department of Botany, University of Maryland, College Park 20742 and National Museum of Natural History, Smithsonian Institution, Washington, DC. 20560. 169 170 GREAT BASIN NATURALIST MEMOIRS No. 2 butes. The family is composed of about 40 genera and approximately 900 species (Law- rence 1951, Melchior 1964, Airy Shaw 1973), with Polygonum L., Rumex L., Eriogonum Michx., Coccoloba P. Br. ex L., Rheum L., and Chorizanthe R. Br. ex Benth. among the larger genera in terms of species numbers. Domestically, the genus Fagopy- rum Mill, is the commercial source of buck- wheat, and leaf petioles of Rheum (rhubarb) are frequently eaten. Antigonon Endl. is an elegant ornamental both in the garden and in nature, although it is more often a weed. A few species of Polygonum, Eriogonum, and Coccoloba are grown for their exotic- properties. The family is usually considered as the only member of a monotypic order, Polyg- onales, which is supposedly related to the Caryophyllales (Takhtajan 1959, 1969, Cronquist 1968, Hutchinson 1969), although some authors still place it with the Caryo- phyllales (Thome 1968, Benson 1974). Re- cently, the relationship with the Caryophyl- lales has been challenged on the basis of pollen data (Nowicke, pers. comm.), and the Polygonales might be better treated as an isolated taxon with no immediate close rela- tives. Polygonaceae has been variously divided into subfamilies (Bentham 1856, Bentham & Hooker 1880, Dammer 1892, Roberty & Vautier 1964), and the differences in opin- ion cannot be resolved here. The one point of near unanimity among all of these au- thors, and others who have treated the Po- lygonaceae, is that Eriogonoideae Benth.2 is the most distinct subfamily of Polygonaceae and can be readily excluded from the re- maining subfamilies. The only serious differ- ence, now largely resolved, has been the relationship of the genus Koenigia L. to the western United States genera, Hollisteria S. Wats., Nemacaulis Nutt., and Lastarriaea Remy in Gay of Eriogoneae Benth., and Pterostegia Fisch. & Meyer of the tribe Pte- rostegeae Torr. & Gray. Bentham and Hooker (1880) proposed that Koenigia was related to these genera, placing all of them in a tribe termed Koenigeae. This was di- rectly contrary to Torrey and Gray (1870) who placed Nemacaulis and Lastarriaea in the tribe Eriogoneae, and Pterostegia in Pte- rostegeae; Koenigia was not even mentioned by Torrey and Gray. Roberty and Vautier (1964) removed Koenigia from Eriogo- noideae, and placed this arctic and sub- arctic genus in the Polygonoideae where it certainly belongs. All further comments in this paper will be restricted to the subfamily Eriogo- noideae. Generic Composition of Eriogonoideae The members of Eriogonoideae are re- stricted to the xeric regions of the New World, with the vast majority of species confined to the western half of central North America from the Tropic of Cancer northward to the fiftieth parallel. In South America, the few species that are known to be native elements in the flora are found in the deserts of northern Chile and scattered parts of adjacent Argentina. The state of California in the United States harbors more species and genera of the subfamily than any other comparable political area; the state also has more endemic genera (five) than any other area. Both Eriogonum and Chorizanthe have a large number of species in California (about 40 percent of Eriogo- num and perhaps 70 percent of Chorizanthe are in the state), and of all the genera of the subfamily, only two, Stenogonum Nutt. and Harfordia Greene & Parry, do not oc- cur in California. Three genera occur in both North and South America. Chorizanthe The authorship of the subfamily name, Eriogonoideae, is here attributed to George Bentham's name "Subordo Eriogoneae" published in deCandolle's Prodromus (14: 5. 1856) based upon Article 18 of the present Code (Stafleu et al. 1972), which states "Names intended as names of families, but pub- lished with their rank denoted by one of the terms order {ordo) or natural order (ordo naturalis) instead of family, are treated as having been published as names of families." Unfortunately, in Article 19, which deals with subfamily, no similar provision is stated. For this reason, some may reasonably ar- gue that the correct authorship of the subfamily is Roberty and Vautier (Boissiera 10: 83. 1964). 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 171 is strictly an annual group in North Ameri- ca where some 40 of the 50 species of the genus are found, but in South America, all but one (C. commissuralis Remy in Gay) of the 10 or so species of the genus are per- ennials, and no species is common to both continents. The genus Lastarriaea is now usually defined to include 2 species (Good- man 1943, Hoover 1966, Munz 1974), L. chilensis Remy in Gay of Chile, and L. co- riacea (Goodman) Hoover of coastal Califor- nia and northern Baja California, Mexico, although the genus has been considered to be monotypic with two variants (Gross 1913, Goodman 1934) or without any differ- ences (Parry 1884, Abrams 1944, Munz & Keck 1959). The third genus, Oxytheca Nutt., has a single species in South America and 8 species are restricted to North Ameri- ca. All of the remaining genera are restrict- ed to North America as are approximately 305 of the 320 species of the subfamily.3 Eriogonoideae is composed of fourteen genera unequally divided into two tribes. The large, typical tribe, Eriogoneae, con- tains twelve genera and about 318 species, with the majority of the species distributed in two genera, Eriogonum (247 species) and Chorizanthe (about 50 species). The least advanced member of the tribe is the genus Eriogonum. Associated with this genus are a series of small, satellite genera which can trace their probable origin to an extant group within Eriogonum as it exists today. Likewise, around Chorizanthe are related genera which probably owe their origins to Chorizanthe, with Chorizanthe itself likely evolved from a now extinct portion of Eriogonum. The second tribe, Pterostegeae, is composed of two monotypic genera which are only superficially related, and whose relationship with Eriogoneae is frank- ly lost. Eriogonum is widespread in North Ameri- ca, ranging from east central Alaska (Welsh 1974) southward to central Mexico, and from the offshore islands of California and Baja California eastward to the Appalachian Mountains of Virginia and West Virginia southward to Florida. In spite of its large size in terms of species numbers, Eriogonum has only three generic synonyms. Eucycla Nutt. (Nuttall 1848a) and Pterogonum H. Gross (1913) are now recognized as sub- genera of Eriogonum (Reveal 1969a, b; Hess & Reveal 1976), while Sanmartinia Buch- inger (1950), a name proposed for Eriogo- num divaricatum Hook. (Reveal & Howell 1976) when it was discovered as an in- troduction into Argentina (Spegazzini 1902, Moreau & Crespo 1969) and thought to represent a distinct species of Eriogonum or a valid genus, is now reduced to synonymy completely. The genus Eriogonum is currently being monographed by myself, but past reviews have been presented by Nuttall (1817), Ben- tham (1836, 1856), Torrey and Gray (1870), Watson (1877), Stokes (1904, 1936), and Re- veal (1969a). Most closely related to Eriogonum is Oxy- theca. This genus of nine species has been reviewed by those who revised the species of Eriogonum (except Nuttall [1817], Ben- tham [1836], and Reveal [1969a]) at least as far as the genus in North America is con- cerned, with both Stokes (1904, 1936) and, indirectly, Roberty and Vautier (1964) in- cluding the species of Oxytheca in Eriogo- num. Critical reviews of the genus have been presented by Jepson (1913), Abrams (1944), and Munz and Keck (1959) in floris- tic studies of the California species where seven of the eight North American species are found. Barbara J. Ertter, a graduate stu- dent at the University of Maryland, is now monographing the genus. One new species has been discovered from California, and, although recognized as unique by Stokes (1904) and by Goodman (in herbaria annota- tions), this San Bernardino Mountains en- demic has not been described. Two generic 'See note added in proof at end of paper. 172 GREAT BASIN NATURALIST MEMOIRS No. 2 segregates have been proposed for species of Oxytheca: Brisegnoa Remy in Gay (1851), a name actually proposed by Remy prior to 1848 when Nuttall described Oxytheca, but whose publication was delayed, and Acan- thoscyphus Small (1898) for a California species of Oxytheca, O. parishii Parry, that differs from most species of the genus in having a multiple-awned, nonlobed in- volucre. Oxytheca luteola Parry is now re- ferred to Goodmania Reveal & Ertter, and O. insignis (Curran) Goodman is placed in Centrostegia Gray ex Benth. in DC. (Good- man 1957). In this latter paper, Goodman informally proposed to divide Oxytheca into two new genera and at the same time sub- merge a part of Oxytheca in Eriogonum. Based upon herbarium annotations, he would have placed O. dendroidea Nutt., O. watsonii Torr. & Gray, O. perfoliata Torr. & Gray, and O. parishii in Eriogonum, O. luteola in a new genus, and referred O. caryophyUoides Parry, O. emarginata Hall, and O. trilohata A. Gray to a second new genus. Goodman did not recognize the South American form of Oxytheca as a dis- tinct species as proposed by Miers (1851), but retained it as a variant of the North American species, O. dendroidea as sugges- ted by Johnston (1929) who proposed var. tonsiflora I. M. Johnst. He and Goodman felt the Chilean and Argentinean plants were a distinct form of O. dendroidea, which they believed also occurred in other areas of South America. Goodman never pub- lished his proposed revision of Oxytheca, and Ertter and I are now investigating the genus. As noted above, one of the species of Oxytheca that Goodman proposed to place in a distinct genus was O. luteola. This sug- gestion has recently been accepted by Re- veal and Ertter (1976b), who proposed the genus Goodmania for this species. Good- mania is restricted to alkaline places, dry lake flats, and similar locations in the south- ern end of the Central Valley of California and elsewhere in the southern part of the state. This monotypic genus seems to be re- lated to both Oxytheca and Gilmania Cov. Somewhat less closely related to Eriogo- num, but still clearly derived from that genus (rather than Chorizanthe), are a series of highly restricted, endemic, western North American genera. Dedeckera Reveal & Howell (1976) is a large perennial shrub re- stricted to a single known location just out- side the northwestern edge of the Death Valley National Monument near Eureka Valley in Inyo County, California. This is the only immediate relative of Eriogonian that is perennial. Stenogonum (Nuttall 1848a) is a genus of two species and is restricted to the Colo- rado Plateau and adjacent regions of Wyoming southward through eastern Utah and adjacent western Colorado into north- ern Arizona and New Mexico. Until recent- ly, this genus was included in Eriogonum, where it had been placed by Hooker (1853) shortly after it was proposed by Nuttall, but it is now considered a valid genus on the basis of its unique involucral construction (Reveal & Howell 1976, Reveal & Ertter 1976a). Gilmania (Coville 1936), another Death Valley region endemic, was originally pro- posed under the generic name of Phyllogo- num Cov. (Coville 1893), but as this name proved to be a later homonym, Coville re- named it for a local Death Valley naturalist, M. French Gilman. Stokes (1904, 1936) maintained the genus as distinct from Eriogonum, although Jones (1903) reduced it to Eriogonum without comment. Roberty and Vautier (1964) placed Gilmania in Eriogonum too; but, unlike Jones, who re- tained the species as distinct, they placed the name in synonymy under Stenogonum salsuginosum Nutt. (which they placed in Eriogonum), an opinion that is totally in- comprehensible. On the Inner Coast Ranges of California is the monotypic genus Hollisteria (Watson 1879). Jones (1908) proposed Chorizanthe floccosa, which proved to be a synonym of H. lanata S. Wats., but it seems unlikely that Jones comprehended the significance of 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 173 his proposal and simply felt the plants rep- resented a species of Chorizanthe, and did not consider that he was reducing Hollis- teria to Chorizanthe. Roberty and Vautier (1964) placed the taxon in Eriogonum, but this concept has not been followed by any- one, and not even Stokes (1904, 1936) felt compelled to reduce Hollisteria to Eriogo- num, although what her opinion might have been regarding its placement in Chorizanthe was never expressed in print. The genus Nemacaulis (Nuttall 1848a, b) is rather widespread in the southwestern United States and extreme northwestern Mexico, with the single species, N. denu- data Nutt, divided into two weakly defined variants. Stokes (1904, 1936) reduced Nema- caulis to Eriogonum, perhaps following the ideas of Curran (1885), who noted the close relationship between N. denudata and E. gossypinum Curran. No one except Roberty and Vautier (1964) has followed this reduc- tion. A second cluster of genera is related to Chorizanthe. Unlike those which have just been reviewed, the satellite genera in this section can be traced to extant sections of Chorizanthe. Chorizanthe itself is a genus of perhaps 50 species, with about 40 species found in west-central North America, and the re- maining 10 or so restricted to northern Chile in South America. All of the species in North America are annuals, while all but one of the South American species are sub- fruticose perennials. The genus has been monographed only by Bentham (1836, 1856), although the North American species have been revised by Torrey and Gray (1870), Watson (1877), Parry (1884), and Goodman (1934). Remy (1851) and Philippi (1864, 1873, 1895) have added species to the South American component of Chori- zanthe. Like Eriogonum, the generic concept of Chorizanthe has changed over the years, with Chorizanthe being defined in both a broad and a strict sense. As the genus is outlined here— and it is done so only in a tentative fashion— a middle-of-the-road ap- proach is proposed. Several segregate gen- era have been proposed from members fre- quently placed in Chorizanthe. These genera are Mucronea Benth., Lastarriaea, Centros- tegia, Acanthogonum Torr., and Eriogonella Goodman. In the following treatment, Mucronea, Lastarriaea, and Centrostegia are recognized as distinct from Chorizanthe, with Acanthogonum and Eriogonella retain- ed in Chorizanthe. The genus Mucronea (Bentham 1836) was described at the same time that Chorizanthe was proposed, and it was retained as a dis- tinct genus by Bentham (1856) in his mon- ograph on Eriogonoideae in deCandolle's Prodromus. Torrey and Gray (1870) reduced Mucronea to Chorizanthe, and their opinion was followed by Bentham and Hooker (1880) a decade later. Goodman (1934) reintroduced Mucronea into the literature when he distinguished it from Chorizanthe in his monograph on the latter genus. How- ever, even with Goodman's study, the genus remained suppressed (Abrams 1944, Munz & Keck 1959) except for Hoover (1970), who recognized the genus in a local flora. As defined here, the genus is considered to have two species, both of which are re- stricted to California. Lastarriaea was proposed by Remy in Gay's Flora Chilena (1851), but it was not associated with the tribe Eriogoneae (Ben- tham 1856) until Torrey and Gray (1870) placed the genus in the tribe. Bentham and Hooker (1880) removed it, Hollisteria, and Nemacaulis, along with Pterostegia, and placed them in the tribe Koenigeae. Except for Dammer (1892), this significant depar- ture has not been followed to any degree. As now defined, Lastarriaea contains two species, one in North America and one in South America (Goodman 1943, Hoover 1966). The genus Centrostegia was published for Asa Gray by Bentham (1856) and consid- ered at the time to be a monotypic genus. In 1870, Torrey and Gray added a second species, but in 1877, Watson reduced the 174 GREAT BASIN NATURALIST MEMOIRS No. 2 genus to Chorizanthe, where it remained until Goodman's revision of Chorizanthe in 1934. In 1957, Goodman revised Centros- tegia, bringing the number of species in the genus to four. One species, Centrostegia in- signis (Curran) A. A. Heller (1910), was originally described as a species of Chori- zanthe by Curran (1885) but placed in Oxy- theca by Goodman (1934) without com- ment. In short, this single, unusual species has been batted around in three separate genera, and it still seems out of place in Centrostegia. As now defined, Centrostegia occurs from Arizona and Utah westward to California, where it is found from Monterey and San Luis Obispo counties southward. Of these three genera, all of which have at one time or another been associated with Chorizanthe, data would now seem to in- dicate that only Mucronea is actually all that close to Chorizanthe, with Centrostegia occupying a position somewhat intermediate between Eriogonum (not Oxytheca) and Chorizanthe, and Lastarriaea well isolated from all of the genera but still closer to the Chorizanthe complex than the Eriogonum complex. Preliminary studies on the Chorizanthe complex have revealed some major areas of investigation for future studies. The most important one is to determine the relation- ship between the northern annuals and the southern perennials in the genus Chori- zanthe. The type of the genus is a South American perennial, C. virgata Benth., and these perennials differ markedly from the annuals. Current plans call for a detailed studv of the South American species, which have not been revised in over 100 years. It is hoped this group of plants will be the subject of a doctoral dissertation. A second area of investigation is whether or not the genus Acanthogonum should be recognized and, if so, what members of Chorizanthe should be placed in it. All of the genera discussed to this point belong to the tribe Eriogoneae. The second tribe of the subfamily is Pterostegeae. This taxon may be characterized by the bisaccate bracts which become enlarged, scarious, and reticulate in fruit, and the consistently op- posite leaves. The genus Pterostegia is a monotypic genus of low, spreading to decumbent an- nual herbs. Described by Fischer and Meyer (1835) from material gathered by the Rus- sian explorers near Fort Ross in California, the genus can be rapidly distinguished from all other members of Polygonaceae by its floral features and fruiting characteristics. A major problem has been how to interpret the bisaccate bracts. Fischer and Meyer completely misunderstood the relationship of the bracts of Pterostegia as they at- tempted to relate these bracts to those of Eriogonum. Bentham (1856) misunderstood the bracts too, attempting to define them as three leaves with a contiguous margin which are expanded into a dorsal wing or crest. Torrey and Gray (1870) stated that the bracts were homologous with the bracts of Nemacaulis, but even this seems most un- likely today, although the concept expressed by them was accepted by Bentham and Hooker (1880). The involucral bracts of Pte- rostegia are two-lobed, enlarged in fruit, and are simply unlike anything found in any genus of Eriogoneae. Pterostegia is a rather variable species which ranges from Oregon southward to northern Baja California and eastward into Utah and Arizona. It does not seem to be divisible into infraspecific elements although Nuttall (1848a) suggested some segregates. The second genus of Pterostegeae, Har- fordia, was proposed by Greene and Parry in a paper published by Parry (1886). The year before, Greene (1885) had described Pterostegia galioides Greene, and, while he placed the species in Pterostegia, it was the first time that good specimens of P. macrop- tera Benth. (Bentham 1844) were found. Bentham's descriptions of his species, pub- lished both in 1844 and 1856, lacked the fine detail, and he was not even sure whether the plants were annuals or per- ennials. It is likely that the lack of adequate material accounted for the long delay in as- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 175 certaining the significant differences be- tween the type species of Pterostegia, P. drymarioides Fisch. & Mev., and P. macrop- tera. Once Harfordia was described, it was immediately accepted, and the genus is now well recognized (Shreves & Wiggins 1964). At present, H. macroptera (Benth.) Greene & Parry is known only from the west coast of central Baja California, Mexico. Origin of Eriogonoideae The origin of Eriogonoideae is unknown. The subfamily is clearly a member of Po- lygonaeceae, for it shares with the other subfamilies of the family a large number of morphological and biochemical similarities, and the subfamily Eriogonoideae cannot be raised to the familial level as proposed by Meisner (1841) without violence to our un- derstanding of families in the Magnolio- phyta. Eriogonoideae has a single, basal, bi- tegmic, crassinucellate ovule similar to that of Polygonum and has the typical trinu- cleate pollen of the family. Still, these are features which are not only typical of Po- lygonaceae, but of Plumbaginaceae and nearly all of the families commonly associ- ated with the Caryophyllales (Cronquist 1968). Eriogonoideae also shares with the other subfamilies of Polvgonaceae the copi- ously laden endospermous seeds and the an- thocyanins pigmentation. The subfamily does differ from the other subfamilies in lacking the distinctly sheathing stipular och- rea of the leaves (although an ochrea is weakly present in some perennial species of Chorizanthe), and the pollen of the sub- family is the least specialized of all sub- families of Polvgonaceae (Nowiche, pers. comm.) suggesting that, as a group, Eriogo- noideae may be a rather primitive member of Polvgonaceae. Equally important in this regard, it may also mean that Eriogo- noideae, as a group, has retained many of the least specialized features of the family due to a lack of modification in organs which have occurred in other taxa. Although the critical similarities between the subfamilies certainly associate these taxa of Polvgonaceae into a distinct family, the place and mode of development of Eriogo- noideae from the rest of Polvgonaceae is now obscured by time and compounded by a lack of a fossil record. No one group of genera outside of the Eriogonoideae can be considered the exact point of origin of the subfamily, and for this reason, the subfamily (or tribe) has long been considered unique within Polvgonaceae (Bentham 1836, Good- man 1934, Boberty & Vautier 1964). Based on preliminary pollen data from extant taxa now available from the work of Dr. Joan Nowicke at the Smithsonian Institution, it seems clear that Eriogonoideae is clearly differentiated from all but the South Ameri- can tropical genus Triplaris Loefl. This is the only genus which has a similar, un- specialized pollen grain (and thus different even from the related American tropical genus Ruprechtia C. A. Meyer), but based on extant data on chromosome numbers, gross morphology, and other anatomical and morphological features (especially in the in- florescence), it seems most unlikely that the tribe Triplarideae C. A. Meyer and the sub- family Eriogonoideae have been connected in any but the most remote fashion. Boberty and Vautier (1964) placed Triplarideae in the subfamily Calligonoideae Boberty & Vautier which they defined as a group of New and Old World genera. Dam- mer (1892) referred the tribe to Coccolo- boideae Dammer, a basically shrubby or ar- borescent taxon about equally divided in the New and Old World. Dammer's suggestion seems more reasonable as he defined the Coccoloboideae to include (using current nomenclature) such genera as Coccoloba, Muehlenbeckia Meisner, and Triplaris— all genera with ruminated endosperm. Unfortu- nately, all genera of Eriogonoideae have a smooth endosperm. Meisner (1856), Ben- tham and Hooker (1880), and Dammer (1892) all placed Eriogonoideae in a posi- tion in their revisions of the Polvgonaceae which would imply that Eriogonoideae is the least specialized of the family. Boberty 176 GREAT BASIN NATURALIST MEMOIRS No. 2 and Vautier placed the subfamily at the end of their treatment; most certainly Roberty and Vautier are correct in their assessment of the placement of the subfamily in the family, for Eriogonoideae is the most ad- vanced member of the extant subfamilies of Polygonaceae and not the least specialized. In Meisner, Bentham and Hooker, and Dammer, interestingly, the Triplarideae was considered the most advanced member of the subfamily Polygonoideae (Meisner and Bentham and Hooker) or Coccoloboideae (Dammer). If this is indeed the case, then it logically can follow that a possible origin of the Eriogonoideae may have been within an ancient taxon that, by definition, might in- clude the basic expression from which the Triplarideae has evolved or in fact was a part. At no time, however, has Triplaris or Ruprechtia played a direct role in the ori- gin of any genus within Eriogonoideae. It is likely that the divergency of Eriogo- noideae from the rest of Polygonaceae has been so fundamental, so sudden, and so suc- cessful, that the new subfamily has com- pletely swamped those groups (or that group) from which it arose. If this divergen- cy is an ancient one, as I suggested some years ago (Reveal 1969b), and occurred at the beginning or slightly before the start of the Tertiary some 65 million years ago, then the loss of such intermediate stages of evolutionary development is to be expected. However, if the origin of the subfamily has been well within the Tertiary, as now seems much more likely, then the loss of the inter- mediate forms is a matter of the explosive success and highly competitive nature of the new form (in this case, the earliest members of Eriogonoideae) as opposed to the rather static parental type (see Stebbins [1974] for a detailed discussion of this type of explosive evolution above the generic- level). As just noted, it now seems more reason- able to assume that Eriogonoideae arose during the Tertiary, and probably during the Oligocene or Miocene epochs (7 to 38 million years ago) when there was a general drying of the climate coupled with the rap- id development and increase of herbaceous angiosperms (Gray 1964, Axelrod 1966, Tid- well et al. 1972). Pollen grains, attributable to Eriogonum, have been found in the Qua- ternary, which began some 7 million years ago (Leopold, pers. comm.). If this as- sumption is correct, then perhaps the sub- family Eriogonoideae had its origin from a subtropical group of New World polyg- onaceous plants near the beginning of the drying period during the Oligocene, which split off into a tropical complex (something like Triplaris) and a northern temperate complex (something like Eriogonum). This point of origin has subsequently been lost, with the extinct relatives of Triplaris and Eriogonum extending and amplifying the differences between the two extremes to a point that now only the mere hint of rela- tionship may be noted in a conservative feature such as pollen morphology. If the theories of Takhtajan (1969) and Stebbins (1974) are correct regarding the differential rates of specialization between tropical and xeric temperate groups, then one may as- sume that the relatives leading to Triplaris have undergone less specialization and dif- ferentiation than the relatives leading to Eriogonum. This would seem to be the case here, especially when one looks at the re- duction of the inflorescence in Eriogonum to a cluster of flowers, the reduced stature of Eriogonum, and the great proliferation of species in the Eriogonoideae when com- pared with the Triplarideae. Over the years I have vainly searched the temperate members of Polygonaceae, and especially those of Asia, for a hint to the origin of Eriogonum. None has been found. Stebbins (1974) has cautioned us to realize that in evolutionary events such as the ori- gin of taxa above the species rank extinc- tion and extensive changes in the distribu- tion and ecology of a taxon may occur which can substantially change one's out- look as to the possible site and point of ori- gin for a given group. It now seems reason- able to look to the New World tropics for a 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 177 point of origin rather than the Old World steppes. Certainly, the pollen data just re- cently reported to me by Nowicke has greatly strengthened this preconceived idea. The origins of the Eriogonoideae very likely have revolved around the reduction of the inflorescence from an extended one (such as in Ruprechtia or Triplaris) to a capitate one, the development of an involucre due to the fusion of subtending bracts on the in- florescence, and the reduction of stature from a shrub or small tree to a subshrub or low shrub. The final step, of course, has been the development of an ability to evolve successfully in a xeric habitat rather than in a mesic, subtropical, or tropical habitat. To my knowledge, none of these steps is extant today. It was proposed by me (Reveal 1969b) that the probable ecological place of origin for Eriogonum was in a xeric site, and that the first forms of the genus were subshrubs or low shrubs. This suggestion has been sec- onded by Stebbins (1974), and there seems to be little reason to alter this opinion. It is important to note that this statement relates to one genus, Eriogonum, and not to the origin of the subfamily. It seems to me that the stages of development leading from the tropical origin of the precursors of the Eriogonoideae to the extant genus Eriogo- num must have taken many different direc- tions, a great deal of time, and undergone many different attempts before arriving at this modern genus. It is now impossible to close that gap, since the history^ of the Eriogonoideae during the Tertiary is un- known. The most generalized form of Eriogonum that exists today is a low, rounded shrub with cauline leaves, cymose inflorescences, small smooth achenes, and an unspecialized flower with monomorphic tepals. These shrubs occur in xeric habitats mainly in the pinyon-juniper woodlands of the Great Ba- sin in Utah and Nevada. Even so, these spe- cies of Eriogonum are highly specialized as all are tetraploids, and no diploid species are known to exist in the genus (Stebbins 1942, Stokes & Stebbins 1955, Reveal 1969b). Therefore, Eriogonum, as it exists today, is a highly evolved group, and no species now exists which could point to the initial element(s) which might have evolved from other, more primitive, subtropical taxa of Polygonaceae. The assumption that Eriogonum is the most basic genus of the subfamily seems reasonable on the basis of morphological considerations, especially in the makeup of the inflorescence and involucre. However, I suspect, that the most ancient extant mem- bers of the subfamily are the perennial spe- cies of Chorizanthe. As shall be discussed below, Chorizanthe likely evolved from Eriogonum, and not the other way around. However, one feature found in these per- ennial species of Chorizanthe seems to hint at their ancientness: they have what can only be considered as weakly defined, fi- brous remains of ochrea. If these species of Chorizanthe should prove to be diploids, this would reinforce their evolutionary sig- nificance. Based upon an examination of the gross morphology of these plants, one must add to the definition of the earliest mem- bers of Eriogonoideae the presence of an ochrea. If the genus Eriogonum underwent its early development in a xeric habitat domi- nated by pinyon-juniper woodlands, then where was such a site in the Miocene or early Pliocene epoches when the genus was undergoing its earliest development? During the Miocene, the Great Basin was dominated by extensive coniferous forest, with the Sierra Nevada to the west about 1000 m in altitude, and thus an ineffective rainshadow (King 1959). It is important to note that these coniferous forests were tem- perate in nature, with the subtropical for- ests of the Oligocene largely pushed to the south. Axelrod (1950) has suggested the ex- istence of two major geofloras, with the Arcto-Tertiary geoflora of hardwood-de- ciduous and conifer forests dominating the Great Basin region, and the Madro-Tertiary geoflora of small-leaved, drought-resistant 178 GREAT BASIN NATURALIST MEMOIRS No. 2 shrubs and trees of the southwestern United States and northwestern Mexico. Axelrod (1958) states that the Madro-Tertiary geo- flora moved northward into the Great Basin in Early Pliocene, but did not entirely re- place the Arcto-Tertiary geoflora. It would seem possible that Eriogonum may have imdergone its early development and differentiation in the Madro-Tertiary geoflora during the Miocene and became well established in the Arcto-Tertiary geo- flora in at least two different expressions: one typified by the subgenus Eucycla (Nutt.) Kuntze in Post & Kuntze (with such species similar to E. microthecum Nutt. or E. corymbosum Benth. in DC.) and the oth- er of members typical of the subgenus Oli- gogonum Nutt. (with such species similar to E. umbellatum Torr. or E. flavum Nutt. in Fras.). Out of the Madro-Tertiary geoflora possibly came such subgenera as Eriogonum and Pterogonum (H. Gross) Reveal which contain such species as E. longifolium Nutt., E. atrorubens Engelm. in Wisliz., and E. alatum Torr. in Sitgr., or their progenitors (Hess & Reveal 1976). Nonetheless, the bas- ic expression of the genus would have be- longed to the subgenus Eucycla, which is basically a taxon of xeric, pygmy coniferous forests. It is also likely that Chorizanthe evolved during this period of time from the subgenus Eucycla, probably when the sub- genus was in the Madro-Tertiary geoflora and before the subgenus underwent its mod- ern-day development of species complexes now typically found in the Great Basin. Evolution within Eriogonoideae If the hypothesis is correct that Eriogo- num is the most primitive extant member of the subfamily Eriogonoideae, then a number of corollaries may be presented. Within Eriogonum itself, if the basic ex- pression of the genus was a low, spreading subshrub or shrub with alternate leaves, cy- mose inflorescences, and unspecialized tep- als. then the subgenus Eucycla was the in- itial expression within the genus. As just noted at the end of the previous section, it is probable that the differentiation of the subgenera Eucycla, Eriogonum, Oligogo- num, and Pterogonum occurred during the Late Miocene or Early Pliocene in the Madro-Tertiary geoflora of northern Mexico and the southwestern United States. Three of these subgenera of Eriogonum are fairly distinct from one another, with no inter- connecting forms. It is felt that while Eriogonum and Pterogonum evolved from Eucycla, these two did not evolve from any extant member of Eucycla. As for Oligogo- num, it is close to Eriogonum and more dis- tantly related to Eucycla, and thus both Oligogonum and Eriogonum may have de- veloped from extinct, primitive members of Eucycla at approximately the same time. Of the remaining subgenera, Clastomyelon Cov. & Morton, Micrantha (Benth.) Reveal, Ganysma (S. Wats.) Greene, and Oregonium (S. Wats.) Greene, all can be traced rather directly to the subgenus Eucycla without any major difficulties. As for Eucycla, it has developed every perennial habit expression of the genus Eriogonum but one, the monocarpic habit of E. alatum of the subgenus Pterogonum. The subshrub or low shrub habit is widely seen in the less specialized members of Eucycla. Such plants are typically seen in the pinyon-juniper (or pygmy) woodlands throughout the western part of central North America today, or essentially the en- tire geographical range of the subgenus which extends from the fiftieth parallel southward to the Tropic of Cancer. Also found in the pygmy woodland zone are sev- eral different kinds of herbaceous perennial expressions belonging to Eucvcla. Unlike the shrubs which tend to be species of wide- spread distribution, the herbaceous per- ennials tend to be more restricted in their range. Some of these species evolved within the zone and have remained while others have extended themselves beyond the con- fines of the zone, and still others, in more recent evolutionary times, have entered the zone from other areas. From the pygmy 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 179 woodland zone, species of Eriogonum have migrated into dry, xeric clay habitats, into grasslands or chaparral habitats, and into al- pine zones and off-shore islands. In all cases, the present members of the subgenus are tetraploid, derived species, and while the majority of morphological expressions are found in the pinyon-juniper belt, the majority of explosive evolution within Erigo- num, and in all of its related genera (except for the most initial phases of Chorizanthe), owe their origins to their survival in ecolog- ical life zones other than the pygmy wood- lands. The origins of both of the predominantly annual subgenera, Ganysma and Oregonium, probably owe their origins to the subgenus Eucycla and probably had an initial differ- entiation in the pinyon-juniper woodlands of the West. Once again the basic expression of these subgenera are generally widespread and found mainly in this habitat. However, unlike portions of Eucycla, it seems— espe- cially in Ganysma— that much of the initial evolutionary development of these sub- genera have been lost over time, because there are widely scattered species of Ga- nysma in other habitats in which the species are obviously primitive but by no means an- cient. The temperate arid regions of the inland portions of North America exhibit the unique combination of selective drought and cold temperatures, conditions which have likely played a major role in the evo- lution of the shrubby habit from which her- baceous and cespitose perennial species could have evolved (Axelrod 1966). The shrubby and subshrubby species of Eriogo- num and Chorizanthe have certainly been subjected to the selective pressures of sea- sonal cold, coupled with enough summer moisture to sustain growth, and the ability to occupy habitats that are protected enough to allow for long-term survival in extended periods of stress. By looking at a pygmy woodlands as the original home of Eriogonum and its first major dichotomy, Chorizanthe, one can understand the variety of habit and morphological expression in this ecological habitat, and the economy in terms of species diversity in this zone. On the whole, the explosive evolution of the modern-day species of Eriogonum and its related genera, and Chorizanthe and its re- lated genera, has been areas in of extreme environmental stress outside the protective (such as they are) confines of the pinyon- juniper belt. Without a doubt, the majority of the genera related to Eriogonum owe their origin to their successful adaptation to a stress condition, mostly selective drought, accompanied by the occupation of ecologi- cal areas on the margins of protective life zones. However, as we shall see, the sub- genus Eucycla has given rise only to the other subgenera of Eriogonum, Chorizanthe, and Dedeckera, but none of the other gen- era. Oxytheca, Stenogonum, Gilmania, Goodmania, Nemacaulis, and perhaps Hol- listeria owe origin to Eriogonum subgenus Ganysma, while Mucronea, Centrostegia, and perhaps Lastarriaea owe their origin to annual species complexes of Chorizanthe. And note, all of these genera (except the primitive members of Chorizanthe) are basi- cally taxa of areas of extreme aridity, and basically adapted to the annual habit (all but Dedeckera). The one major adaptation which dis- tinguishes Chorizanthe from Eriogonum is a combination of the production of an awned involucre and the reduction in the number of flowers per involucre. I believe the pro- duction of an awned involucre has occurred several times in the history of the sub- family, much as the total loss or reduction of an involucre has occurred several differ- ent times and places in the taxon. The key to understanding the origin and evolution of Chorizanthe lies in the unstu- died South American perennial species. An examination of available herbarium material seems to point the origin of these perennials to the subgenus Eucycla of Eriogonum, a theory which seems reasonable if Eriogonum is, as I suspect, the basic element of the subfamily. There are, however, some diffi- 180 GREAT BASIN NATURALIST MEMOIRS No. 2 culties which in theory can be excused but need to be mentioned. If Chorizanthe evolved from Eriogonum subgenus Eucycla, it did not evolve from any extant group of the subgenus. One can account for the subshrubby habit of the per- ennial Chorizanthe as having come from Eu- cycla. Even the hooked, awned condition of the involucre could be traced to the sub- genus as several extant species of Eucycla have long, sharply acute involucral lobes which, while not awned, could point to a stage in the development of the awned con- dition. The narrow, essentially basal leaves of Chorizanthe can be traced to Eucycla, as can the congested, cymose inflorescence. Two major drawbacks exist. One is the straight embryo of Chorizanthe (Goodman 1934), whereas all species of Eucycla have a curved embryo (Reveal 1969a, b). The sec- ond is the six-lobed involucre of Chori- zanthe, while the majority of species in Eu- cycla are five-lobed. The critical hint here, I believe, is the presence of the remains of the ochrea in some species of South American Chori- zanthe. As Grant (1971) has noted, a given character may or may not be selected for or against, and thus, while the direction of the subfamily Eriogonoideae has been to get rid of the ochrea, at sometime in its history of divergency from the rest of Polygonaceae it must have possessed this feature. If, as I sus- pect, the perennial species of Chorizanthe are the most ancient extant members of the subfamily, then it would follow that these plants would exhibit some of the more primitive features of the subfamily and pro- vide helpful keys to its origin. By the same token, while I accept Eriogonum as the bas- ic expression of the subfamily, and Chori- zanthe as a derived element, one need not look further than extant and derived mem- bers of Eriogonum to find all unique fea- tures of the South American perennials ex- cept the ochrea. Thus, if Chorizanthe evolved as a preliminary expression from Eucycla, as did the subgenera Eriogonum or Oligogonum, then suddenly we find species of Eriogonum '•'.n a straight embryo and a six-lobed involucre. The genus Oxutheca, which can trace its immediate origin to Eriogonum subgenus Ganysma, has awned involucres, and the reduction in the number of flowers per involucre can be seen in sev- eral different groups of Eriogonum, although admittedly this feature is almost entirely re- stricted to annual species. The next critical step in this discussion is how did the perennial members of Chori- zanthe get to South America while Eriogo- num did not, and if Chorizanthe evolved from Eriogonum subgenus Eucycla in North America, why are there no perennial spe- cies of Chorizanthe in North America? The first part of this question can be eas- ily answered. The only members of Eriogo- noideae in South America are those which have a distinctly awned, or hooked, in- volucral lobe. As Stebbins (1974) has point- ed out, such an adaptation can be a success- ful means of long-distance dispersal. Thus Eriogonum (with the exception of E. diva- ricatum, an annual species which was found as a waif in eastern Argentina) is perhaps lacking from South America due to the ab- sence of an awned involucre. The second part of this question, why the perennial spe- cies of Chorizanthe are missing from North America, is much more difficult. Two options exist about the existence of perennials in South America and their lack in North America, and a third option can be proposed on the basis of either of the first two options if future studies should make such an option necessary from a tax- onomic point-of-view. The first two options are closely inter- twined and deal with the actual origin of the perennial species in South America and the annual species in North America. As- suming the idea that Chorizanthe evolved as a perennial group from Eriogonum in North America, then it had to have migrated to South America as a perennial and become established as a perennial. The South Amer- ican populations, I believe, have remained essentially unchanged since their (or its) in- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 181 itial introduction with some speciation oc- curring there within rather limited parame- ters. One hint that this is so is that all of the perennial Chorizanthe species fall within extremely narrow limits morphologically, and, while several species (close to 25) have been described, the actual number of valid species seems to be much less than that. The one annual species in South America is apparently a much more recent in- troduction than the perennial species be- cause it is closely related to the single most widespread annual species in North Ameri- ca (Goodman 1934). The first option states that Chorizanthe evolved in North America and migrated as a perennial to South America as a single in- troduction, with the North American per- ennials gradually being replaced by annual species. In South America, the perennial species were subjected to little direct selec- tion pressure, while in North America the perennial members of Chorizanthe were subject to intense pressures from the rapidly evolving and highly competitive, closely re- lated genus Eriogonum. In order to survive and compete against Eriogoniim, which, I feel, was rapidly adopting the annual habit, Chorizanthe also had to change if this hy- pothesis is feasible. The second option is that the perennial species of Chorizanthe in North America be- came extinct, while the South American species remained. The annual habit then de- veloped in Chile, and only the annual spe- cies were introduced into North America. Raven (1963) has noted that, while the ma- jority of species probably migrated north to south, some certainly went from south to north. Once in North America, the annual species underwent active adaptive radiation similar to that observed in such annual groups of Eriogoniim as the subgenera Ga- nysma and Oregonium. The third option states that the South American perennials represent a genus of plants distinct from the North American (and one South American) annuals. If this is so, then the name Chorizanthe would be ap- plied to the South American perennials, while the annual species would be called Acanthogonum, or, if that genus proves dis- tinct, Eriogonella. This option takes on added significance if the following scenario should prove correct after careful system- atic studies. If indeed Chorizanthe evolved as a perennial and migrated southward, and the northern element became extinct, did the annual species evolve prior to the ex- tinction of the perennial group or did the annual species begin from a whole new series of events? In option one, I have ac- cepted the first part of this question, but if the second were the case, then it will be impossible to retain the North American an- nuals in the genus Chorizanthe. The recently discovered Dedeckera eu- rekensis probably evolved from the sub- genus Eucycla of Eriogonum, and most likely from the section Corymbosa. Its ori- gin is likely most recent. It differs from all other members of the subfamily in having a head of subsessile or sessile flowers, borne on a slender peduncle and subtended by two to five foliaceous bracts, and a single, short-pedicellate axillary flower at the base of each peduncle. It differs from Eriogonum in lacking an involucral tube. This mono- typic genus is known only from a single site where about 200 individual plants are found. It is likely that the genus evolved in place within recent history and, while its range has expanded and decreased through- out its brief history, it is unlikely that the plant has been beyond the restrictive eco- logical confines of the Death Valley region of eastern California. The pubescence of Dedeckera is similar to that of Eriogonum intrafractum Cov. & Morton, another Death Valley endemic, which is the only representative of the sub- genus Clastomyelon and a few other mem- bers of Eriogonum. Of all the subgenera of Eriogonum, Clastomyelon is the most dis- tinctive on pure morphological grounds in that the stems are broken into a series of ringlike segments, the numerous flowers rupture the involucral tube into irregular 182 GREAT BASIN NATURALIST MEMOIRS No. 2 segments, and the bractlets are foliaceous at least in part. While it is possible to trace the origin of E. intrafractum to the sub- genus Eucycla, where D. eurekensis also evolved from, both are amazingly distinct, with D. eurekensis significantly more so than E. intrafractum. It is interesting that in the Death Valley area, where speciation has been rather spectacular (Stebbins & Major 1965), Polygonaceae should be blessed with so many different expressions. Much like Gihnania, which will be discussed below, Dedeckera and E. intrafractum have come about in recent times, influenced by the en- vironmentally profound selection pressures of the area. The remaining satellite genera related to Eriogonum evolved from the subgenus Ga- nysma. Oxytheca is being studied currently by Ertter and me to determine the exact make- up of this genus. We have excluded O. lu- teola, placing it in a new genus, Good- mania. The remaining nine species, however, may or may not be all related. Oxytheca dendroidea, O. watsonii, and the South American plants are related to Eriogonum spergulinum A. Gray, and, based on this close morphological similarity, Goodman (in herbaria) has placed these plants in Eriogonum. Oxytheca perfoliate! probably belongs to this complex of species (Goodman would have placed the taxon in Eriogonum), but it is morphologically dis- tant from the other members. As for O. par- ishii and an undescribed taxon from the San Bernardino Mountains of California, they present a problem. Goodman (in herbaria) would have placed these in Eriogonum, but both seem more closely related to E. api- culatum S. Wats, and E. parishii S. Wats, than E. spergulinum. Small (1898) placed O. parishii in a monotypic genus, Acanthos- cyphus. It is possible that Acanthoscyphus should be recognized if it can be shown that O. parishii and its related taxon are distinct from that group of Oxytheca species typified by O. dendroidea. This complex of species (excluding the Oxytheca parishii complex for a moment) seems to have developed in the pygmy woodlands of the Great Basin and, in par- ticular, along the western edge of the Great Basin. This is a complex of volcanic sandy soils that are widely scattered. I suspect the group evolved during the Quaternary (prob- ably the Pleistocene), with the introduction of the South American phase in recent geo- logical time (see Raven 1963). Goodman (in herbaria) placed Oxytheca caryophylloides, O. emarginata, and O. trilo- bata in a new genus. Our preliminary stud- ies of these species seem to indicate that they too developed from Eriogonum api- culatum, E. parishii complex, and perhaps one should consider if these species too ought not to be referred to Acanthoscyphus. These three species differ from O. parishii and its undescribed relative in having a five-lobed involucre instead of the nonlobed tube with 4 to 30 long bristled awns. All of these plants are found in the granitic moun- tains of southern California and northern Baja California and occur in approximately the same type of ecological niche. I have come to look upon Oxytheca par- ishii, O. caryophylloides, O. emarginata, and O. trilohata as a group that has evolved in the Pleistocene in the mountainous region of southern California and adjacent Mexico as the result of sudden and explosive evolu- tionary changes in the gene makeup of the rapidly developing annual species of both Eriogonum and Chorizanthe. This is not to say that Chorizanthe played a direct role in the development of these species, but one should remember that Centrostegia (and in particular C. insignis) probably developed at the same time, and this genus is similar to Chorizanthe. Thus, I suspect, a whole series of rapid changes were in the process at this time in a small portion of Eriogo- num, which possibly resulted in this group of Oxytheca, Centrostegia, and perhaps (if option three is correct) the annual species of Chorizanthe as well. If this conclusion should prove correct with regards to the species now placed in Oxytheca, then Good- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 183 man would have been proved correct, and these species would have to be placed in a different genus. Stenogotium is a step-child in this group of satellite genera. It is closely related to Eriogonum, differing mainly in the construc- tion of the involucral bracts. While in Eriogonum the involucre is distinctly tubu- lar, the involucre of Stenogonum is com- posed of two whorls of three lobes. That this condition is possible within a genus clearly and closely related to Eriogonum is an important step, because it does demon- strate the potential for a six-lobed or three- lobed involucre as found in the annual spe- cies of Chorizanthe. Stenogonum evolved from the Eriogonum inflatum Torr. & Frem. complex of the sub- genus Ganysma. It is thought that the origin of this clay-inhabiting genus is relatively re- cent and has evolved to a point about on the par with the degree of divergence seen in Oxytheca. Stenogonum is an annual which has evolved from a "hot desert" com- plex on the Colorado Plateau, which is an area somewhat intermediate between the Mojave Desert and the Great Basin in terms of phvsiological stress. Eriogonum inflatum var. inflatum is found on the Plateau, but the more common phase is not the per- ennial var. inflatum, but the annual var. fusiforme (Small) Reveal. Likewise, var. in- flatum is usually found in rocky places above the clay hills and flats, while var. fusiforme is typical of the clay sites. Thus it is that the genus Stenogonum has evolved by successfully occupying the clay habitat that, for the most part, members of the £. inflatum complex cannot enter. Two genera are difficult to directly asso- ciate with Eriogonum, and both perhaps have recently evolved in the subfamily. They are Goodmania and Gilmania. The two seem to be related, as both are pros- trate to low-spreading annuals with pub- escent yellow flowers, cauline leaves, and small, smooth achenes. Goodmania has in- volucral bracts which subtend each cluster of flowers and act as a protective involucre. hi Gilmania, all involucral bracts are lack- ing, but the three foliaceous leaves, when the plants are immature (but that particular branch is in full flower), are held close to- gether by the shortened internodes so that each cluster of flowers is positioned above the lower whorl of three leaves so that the flowers are protected both by these leaves and the whorl of upper leaves as well. In this condition, the long pedicels extend the ripened flowers beyond the protective con- fines of the three leaves so that pollination may occur. In this fashion, the flowers of Gilmania are better protected from the ele- ments than those of Goodmania. Goodmania could possibly be traced to Oxytheca, but I think not. True, it has an awned involucral bract, but, in fact, these bracts are just that and they are not ar- ranged into a distinct tube. The individual bracts can be separated from each other without disruption of tissue on an adjacent bract. There is one bract that is longer than the other four bracts which is unlike any species of Oxytheca, but is a condition that is seen in some species of annual Chori- zanthe. The flowers of G. luteola are yel- low, and no species of Oxytheca has yellow flowers, and the plants of this species are glabrous and bright green while those of Oxytheca are glandular (at least in part) and usually reddish or grayish in color. As I look about the subfamily, I see a possible close relationship with the subgenus Orego- nium for this genus and Gilmania, and in particular Eriogonum divaricatum. The sub- genus Oregonium underwent a major up- heaval in the hot, dry foothills of western California, but E. divaricatum, E. pub- erulum S. Wats., and other similar species are more typical of the Great Basin. Thus, while this group of species of Eriogonum may hint as a possible place of origin for Goodmania and Gilmania, the group seems unsatisfactory, and no extant subfamily of Eriogonum really reveals a logical place of their origin. As noted above, these two genera seem to be recently evolved genera. Goodmania is 184 GREAT BASIN NATURALIST MEMOIRS No. 2 usually found on the plains of old dry lake beds in areas which were covered by water during recent glacial periods. Gilmania oc- curs on the lower rim of Death Valley on alkaline soils near sea level, and thus in areas that were covered by water less than 50,000 years ago. It is likely, therefore, that both genera underwent their evolutionary development at approximately the same time, taking advantage of the same type of opening environment niche. The origin of Gilmania is somewhat more difficult to postulate than that of Good- mania. Cauline leaves in Eriogonum annuals are infrequent, and when present are rarely arranged in a pattern similar to that of Gil- mania, nor are they like the leaves of Good- mania. In Goodmania the leaves are two and opposite, varying from laminar at the lower nodes to acicular at the upper nodes. In Gilmania, the leaves are in threes, with two of the leaves opposite, and the third opposite the next branch; all of the blades are laminar. It seems unlikely that both Gil- mania and Goodmania evolved from pre- cisely the same element within Eriogonum, but they probably did arise within the same subgenus. I strongly suspect that the selec- tive evolutionary pressures have been much greater on Gilmania than Goodmania, thus accounting for the great degree of demarca- tion of Gilmania. Curran (1885) was the first to call atten- tion to the close relationship between Eriogonum and Nemacaulis. Nemacaulis is similar to E. gossypinum in that both have copious bractlets and hairs surrounding and protecting the flowers; in Eriogonum the tubular involucre is broadly campanulate, but in Nemacaulis the involucre is lacking and replaced by subtending bracts. Beyond this, the two taxa are notably distinct. Still, it seems likely that the origin of Nemacaulis can be traced to Eriogonum subgenus Ga- nysma and, in particular, the section of Ga- nysma which contains E. gossypinum. Looking upon Nemacaulis as a recent de- rivation from Eriogonum, it seems to have undergone rapid development in the hot deserts of southern California and adjacent Mexico, occupying a position on the south- ern geographical edge of Eriogonum section Ganysma. I suspect the degree of difference between Eriogonum and Nemacaulis is on the magnitude of that exhibited by Eriogo- num and Oxytheca. The genus Hollisteria is a most difficult genus to trace back to its possible point of origin. It is a prostrate, spreading annual with two sessile, yellow, woolly flowers sub- tended by three slightly united involucral bracts. In some respects, Hollisteria is inter- mediate between Eriogonum and Chori- zanthe. It differs from both in lacking a dis- tinct involucral tube, but it is two-flowered and thus similar to Centrostegia and has acerose tips on the bracts similar to those on Goodmania. It probably did not evolve from an unknown perennial group as pro- posed by Stebbins (1974) but more likely developed from an annual complex. I would like to say that Hollisteria could have evolved from either Eriogonum sub- genus Ganysma or Oregonium, but no ex- tant group in either subgenus can really point the way. I have tried to place the genus near Chorizanthe, but still no one group of that genus really is helpful. In some respects, I have tried to fit it into a pigeonhole between what Goodman termed Eriogonella (C. membranacea Benth.) and Centrostegia because here one can find a combination of three-lobed involucres, yel- lowish flowers, and a spreading annual habit. Still, one compelling bit of evidence that wrenches this entire scene is the nature of the pollen grain. As Nowicke (pers. comm.) has recently shown, the pollen grains of Hollisteria and Lastarriaea are es- sentially the same, and unlike any other genus of Eriogonoideae. It is possible that both Hollisteria and Lastarriaea evolved from an extinct, independent complex of annual species. One part of the complex close to Eriogonum could have given rise to Hollisteria while another part of the com- plex close to Chorizanthe gave rise to Last- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 185 As one might suspect from the foregoing discussion, the genus Lastarriaea is also somewhat intermediate between Eriogonwn and the annual species of Cfiorizanthe, but closer to the latter than the former. Last- arriaea is a low, often spreading annual without a distinctly tubular involucre, ace- rose bracts, and whitish, glabrous, co- riaceous tepals. Unlike all of the genera dis- cussed to this point (with the exception of some species of Chorizanthe and Netna- caulis), Lastarriaea has only three anthers per flower instead of the usual nine. And unlike Hollisteria, which is an inland species of the Inner Coast Ranges of California, Lastarriaea is a coastal genus found in both North and South America. Goodman (1934) placed Lastarriaea in Chorizanthe in the least specialized section of the genus, and, while there are some su- perficial similarities between the section Suffrutices Benth. and Lastarriaea to the point that perhaps it evolved from this sec- tion, I doubt that its point of origin can be traced to any extant section of Chorizanthe. As noted above, a more likely situation is that Lastarriaea developed early in the evo- lution of the annual species of Chorizanthe (or less likely, Eriogonum). I strongly sus- pect that Lastarriaea became well estab- lished in North America, and that it, the one annual species of Chorizanthe, and Oxij- theca all migrated to South America at ap- proximately the same time as hitchhikers on animals, probably during the Late Pliocene (Raven 1963). The differences between the North and South American elements are not strongly expressed morphologically in these annual species, although a strong difference does not necessarily have to be expressed (Grants 1967). Mucronea is clearly derived from the an- nual species of Chorizanthe. This genus has a distinctly tubular involucre like Chori- zanthe, but it and Centrostegia differ in having three-lobed bracts instead of the typ- ically entire bracts of Chorizanthe. The con- spicuous bracts of Mucronea are united and distinct, and in this feature the genus is sim- ilar to Eriogonum and Oxytheca, especially O. perfoliata. Mucronea is distinct from Centrostegia and Oxytheca in having straight cotyledons (Goodman 1934), but is similar to Chorizanthe in this regard. I sus- pect that Mucronea is a rather recent in- novation within the Chorizanthe complex. The genus Acanthogonum was recognized as a distinct genus by Goodman (1955), but I am still somewhat reluctant to recognize it. When Torrey (1857) described Acan- thogonum, he placed a single species, A. rig- idum, in the genus. In 1858, Torrey ques- tionably added a second species, A. corrugatum, noting that this species was "al- most intermediate between Acanthogonum and Chorizanthe." Torrey and Gray (1870) reduced both species to Chorizanthe and added to the complex C. polygonoides and C. watsonii. Goodman (1934) defined Acan- thogonum to include A. rigidum and A. po- lygonoides (Torr. & Gray) Goodman, and these two species were retained in the genus in his 1955 review. Basically Good- man maintained the genus on the basis of the curved cotyledons, but, as I am retain- ing Eriogonella in Chorizanthe, which was established (in part) on its curved cotyle- dons, I cannot very well recognize Acan- thogonum because of this feature. For now, at least, the relationship between C. polyg- onoides and C. corrugata, C. watsonii, and C. orcuttiana Parry seems too close to allow for a distinct genus to be established. Centrostegia is a most difficult and di- verse assemblage of species. As defined by Goodman (1957), the genus consists of four species, three of which, C. leptoceras A. Gray, C. thurberi Gray ex Benth. in DC, and C. vortriedei (Brandeg.) Goodman, form one distinct element within the genus, but C. insignis is decidedly aberrant although even C. vortriedei is somewhat strange with- in Centrostegia. I am inclined to restrict Centrostegia to C. thurberi and C. leptoceras but am lost when it comes to C. vortriedei, and feel C. insignis should probably go into a distinct genus. This latter species is cer- tainly most closely related to Oxytheca, 186 GREAT BASIN NATURALIST MEMOIRS No. 2 where Goodman (1934) placed it at one time. As for C. thurberi and C. leptoceras, one might look for an origin somewhat in- termediate between Eriogonum and Chori- zanthe. Until these species can be carefully studied, especially cytologically, little can be expressed about their relationships. I have little faith in the one unifying charac- ter, which is the three-lobed bract, and would like to place more emphasis on the involucral, floral, and vegetative features of these plants. Up to this point, the discussion has cen- tered on the tribe Eriogoneae, which makes up the vast bulk of Eriogonoideae. The oth- er tribe of the subfamily, Pterostegeae, con- tains only two monotypic genera. Time and evolution have largely destroyed the inter- connecting links between the two tribes so that it is impossible to say what, if any, role Eriogoneae might have played in the evolu- tion of Pterostegeae, or the other way around for that matter. The inflated bracts of the fruiting specimens are unseen in Eriogoneae, and the consistently opposite leaves are rare. I suspect that the two tribes are well separated now by time and events. Until Parry's (1886) paper in which the genus Harfordia was described, the true na- ture of this narrowly restricted shrub was unknown. Bentham (1844) had placed the perennial in the genus Pterostegia not know- ing if his species, P. macroptera, was a shrub or not. It remained there until 1886. Only Roberty and Vautier (1964) reduced Har- fordia to Pterostegia. In spite of this, there is little reason to closely associate Harfordia with Pterostegia except in the feature of the fruiting bracts and opposite leaves. Hutchinson (1926, 1959, 1969) was a firm believer in the concept that certain families of flowering plants were fundamentally her- baceous or woody. Polygonaceae, in his view, was basically a herbaceous group in which the woody, or shrubby, condition was a secondary state. There is something to say about this point, although it may seem con- trary to the usual dicta (Bessey 1915). The largest forms of Eriogonum, for example, are highly derived forms from low, sub- shrubby or shrubby groups. Eriogonum aus- trinum (S. Stokes) Reveal is an annual spe- cies that will form perennial individuals, and this is a condition that will be seen in other species as well. The reason to bring this controversial subject up is Harfordia, the perennial, versus Pterostegia, the annual. It is possible that Harfordia represents the residue of an ancient series of events in a perennial line of evolution from which, at some time in the past, the ancestral fore- runners of Pterostegia evolved. This is the reasonable approach. Another which cannot be totally ignored is that Harfordia is a sec- ondarily evolved perennial which developed from an annual group in order to survive in the extreme stress of long-term drought as- sociated with the environment of central Baja California. Pterostegia is basically a mesic species, and I do not propose to im- ply that Harfordia evolved from Pterostegia, but anatomical and cytological studies may be helpful in unraveling this question. Summary The subfamily Eriogonoideae is divided into two tribes, Eriogoneae and Pteros- tegeae, which are somewhat atypical mem- bers of the Polygonaceae. The subfamily is restricted to the more xeric areas of central North America and western South America. The basic extant expression in the subfamily is Eriogonum, whose ancestral roots can probably be traced to the tropical or sub- tropical members of the family. Chorizanthe was a major side-shoot from Eriogonum, and from these two fundamental genera have evolved a series of small, usually closely re- lated genera. Eriogonum and the perennial forms of Chorizanthe probably developed in the pygmy woodlands of the Madro-Tertiary geoflora, with the related genera evolving mainly in the more xeric, hot deserts at ele- vations lower than those in which Eriogo- num is typically found. The largest and most diverse genus is Eriogonum, both in terms of numbers of 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 187 species and in expressions. Chorizanthe is the next largest, but the degree of mor- phological divergency in this genus is not as great as in Eriogonum. As for the remain- ing, smaller genera, each attempts to fill an available morphological gap or ecological niche, and for the most part, each is suc- cessful. Much work remains to be done on the subfamily. The South American species of Chorizanthe must be studied in the field. The nature of the relationship between the South American perennials and the North American annuals of this genus must be de- termined, and then, if the two should prove distinct, we must decide by what name the annuals should be called. Field studies are now critically needed so that anatomical and cytological material can be gathered, and perhaps greenhouse investigations made. A series of monographic studies are now in progress, mainly on the genus Eriogonum and its immediate relatives. In time, these studies must be expanded beyond the alpha taxonomic level where they are now. This will be a continuing challenge to anyone wishing to travel, study, and investigate one of the world's most unique groups of flowering plants. Acknowledgments The present paper was prepared at Brig- ham Young University where I was on sab- batical leave from the University of, Mary- land the first six months of 1976. To the various authorities at these two institutions, and in particular Dr. Hugh D. Sisler of the University of Maryland and Dr. Bruce N. Smith of Brigham Young University, chair- men of the Department of Botany at their respective universities, I am most grateful for the opportunity to be free for the one semester to work on the proposed mon- ograph on Eriogonum. The field and herba- rium studies which have formed the basis for this paper have been supported by Na- tional Science Foundation grants GB-22645 and BMS75- 13063. I wish to thank my col- league, Dr. C. Rose Broome, of the Univer- sity of Maryland, for sharing with me her ideas and thoughts on the complex problems of evolution and speciation as they might apply to Eriogonoideae. Note Added in Proof Since this manuscript was completed in June 1976, the revision on Oxytheca has been completed [Ertter, B. J. 1977. A revi- sion of the genus Oxytheca (Polygonaceae). Unpublished Master's Thesis, University of Maryland Library, College Park] and a number of minor changes must be ap- pended. Ertter found that Oxytheca consists of seven (not nine) species, with the "South American form" being merely a subspecies of O. dendroidea, and the "new species" from the San Bernardino Mountains of Cali- fornia a variant of O. parishii. Ertter con- curs that O. dendroidea and O. watsonii are related to Eriogonum spergidinum, and she has shown conclusively that O. perfoliata is clearly related to O. dendroidea. In fact, O. watsonii, a rare species of west-central Ne- vada, is intermediate between O. dendroidea and O. perfoliata in many respects accord- ing to Ertter. Ertter also concurs that Oxytheca parishii is most closely related to Erigonum apicula- tum and E. parishii, but in doing so called attention to an error in my own work on Eriogonum (Reveal 1969a). In my revision of Eriogonum I placed E. spergidinum and E. apiculatum in widely separated sections of the subgenus Ganysma. Ertter has shown that these two species complexes are much more closely related than I had thought. In the present paper I raised the question that if O. dendroidea and its allies arose from the E. spergidinum complex (in one part of Ganysma), and O. parishii arose from E. apiculatum (in another part of Ganysma), then perhaps the genus Acanthoscyphus should be recognized. Such a situation now is unnecessary. The relationship between Oxytheca par- ishii and the remaining members of the 188 GREAT BASIN NATURALIST MEMOIRS No. 2 genus in southern California (O. caryophyl- loides, O. trilobata, and O. emarginata) is still tenuous. Ertter has shown, however, that O. parishii is more similar to the O. dendroidea complex than it is to the O. caryophylloides. She has proposed to place O. dendroidea, O. watsonii, and O. per- foliate! in their own typical section, with O. parishii in a monotypic section. As for the other three species, these are going into a third section, a taxon somewhat removed from the other sections. I still believe that the southern California elements evolved as a group in the moun- tainous regions of southern California dur- ing the Pleistocene, but I now feel that this development came not from isolated ele- ments with Eriogonum but from a broadly connected group of annual species all be- longing to this one genus which were un- dergoing collectively rapid evolution (see Raven and Axelrod discussed below). As for the troublesome Centrostegia in- signis, recent conversations with Goodman confirm the supposition that this species is seriously out of place in Chorizanthe (Good- man 1934), Oxytheca (Ertter 1977, cited above), and even Centrostegia (Goodman, pers. comm.) and that it most likely will have to be placed in its own monotypic genus. This question is now being explored. Nowicke's pollen work discussed above has now been published [Nowicke, J. W., and J. J. Skvarla. 1977. Pollen morphology and the relationship of the Plumbaginaceae, Polygonaceae, and Primulaceae to the order Centrospermae. Smithsonian Contr. Bot. 37: 1-64.]. Lastly, my manuscript has been used by Drs. Peter H. Raven and Daniel I. Axelrod in a book entitled Origin and Relationships of the California Flora published by the University of California Press. Based upon my review of their manuscript (which should be published at approximately the same time as this volume), their work will nicely complement the present volume. Literature Cited Abrams, L. 1944. Illustrated flora of the Pacific- states. Vol. II. Stanford University Press, Stan- ford, California. Airy Shaw, H. K. (Rev.) 1973. A dictionary of the flowering plants and ferns. 8th ed. Cambridge University Press, Cambridge. Axelrod, D. I. 1950. Evolution of desert vegetation in western North America. Publ. Carnegie Inst. Wash. 590: 217-306. 1958. Evolution of the Madro-Tertiary geo- flora. Bot. Rev. 24: 433-509. 1966. Origin of deciduous and evergreen habits in temperate forests. Evolution 20: 1-15. Benson, L. 1974. Caryophyllales. Encyclop. Brit- annica 3: 973-976. Bentham, G. 1836. On the Eriogoneae, a tribe of the order Polygonaceae. Trans. Linn. Soc, Bot. 17: 401-420. 1844. The botany of the voyage of HMS Sul- phur, under the command of captain Sir Ed- ward Belcher . . . during the years 1836-1842. Smith, Elder & Co., London. 1856. Eriogonum. In: A. deCandolle, Pro- dromous systematis naturalis regni vegetabilis 14: 5-23. Bentham, C, and J. D. Hooker. 1880. Genera plantarum. Vol. Ill, part 1. Reeve and Co., London. Bessey, C. E. 1915. The phylogenetic taxonomy of flowering plants. Ann. Missouri Bot. Gard. 2: 109-164. Buchinger, M. 1950. Sanmartinia, nuevo genero Pa- togonico de poligonaceas. Comun. Inst. Nac. In- vest. Ci. Nat., Ser. Ci. Bot. 1(4): 1-13. Coville, F. V. 1893. Botany of the Death Valley expedition. Contr. U.S. Natl. Herb. 4: 1-363. 1936. Gilmania, a new name for Phyllogo- num. J. Wash. Acad. Sci. 26: 209-213. Cronquist, A. 1968. The evolution and classifica- tion of flowering plants. Houghton Mifflin Co., Boston. Curran, M. K. 1885. Classification of the Erio- goneae as affected by some connecting forms. Bull. Calif. Acad. Sci. 1: 272-275. Dammer, U. 1892. Polygonaceae. In: A. Engler, and K. Prantl, Die natiirlichen Pflanzenfamilies 3(70): 1-36. Fischer, F. C. L., and C. A. Meyer. 1835. Index seminum, quae hortus botanicus and imperialis Petropolitanus pro mutua commutatione offert. St. Petersburg. Goodman, G. J. 1934. A revision of the North American species of the genus Chorizanthe. Ann. Missouri Bot. Gard. 21: 1-98. 1943. Notes on Chorizanthe. Leafl. W. Bot. 3: 230. 1955. The genus Acanthogonwn, tribe Erio- goneae. Leafl. W. Bot. 7: 234-236. 1957. The genus Centrostegia, tribe Erio- goneae. Leafl. W. Bot. 8: 125-128. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 189 Grant, V. 1967. The selective origin of in- compatibility barriers in the plant genus Gilia. Amer. Naturalist 100: 99-118. 1971. Plant speciation. Columbia University Press, New York. Gray, J. 1964. Northwest American Tertiary pa- lynology: the emerging picture. Pacific Sci. Congr. Ser. 10: 21-30. Greene, E. L. 1885. Studies in the botany of Cali- fornia and parts adjacent. Bull. Calif. Acad. Sci. 4: 179-228, 276-282. Gross, H. 1913. Beitrage zur Kenntnis de Polyg- onaceen. Bot. Jahrb Syst. 49: 234-339. Heller, A. A. 1910. Catalogue of North American plants. 3d ed. Published by the author, Los Gatos. Hess, W. J., and J. L. Reveal. 1976. A revision of Eriogonum (Polygonaceae), subgenus Pterogo- num. Great Basin Nat. 36: 281-333. Hooker, J. D. 1853. Catalogue of Mr. Geyer's col- lection of plants gathered in the upper Mis- souri. Hooker's J. Bot. Kew Gard. Misc. 5: 257- 265. Hoover, R. F. 1966. Miscellaneous new names for California plants. Leafl. W. Bot. 10: 337-350. 1970. The vascular plants of San Luis Obispo County, California. University of California Press, Berkeley. Hutchinson, J. 1926. The families of flowering plants. I. Dicotyledons. Macmillan & Co., Lon- don. 1959. The families of flowering plants. Ox- ford University Press, New York. 1969. Evolution and phylogeny of flowering plants: Dicotyledons; fact and theory. Academic Press, London. Jepson, W. L. 1913. Eriogonum. A flora of Califor- nia 1: 376-428. Johnston, I. M. 1929. A collection of plants from the high cordilleras of northwestern San Juan. Physis (Buenos Aires) 9: 297-326. Jones, M. E. 1903. Eriogonum. Contr. W. Bot. 11: 4-18. 1908. Contribution to western botany. Contr. W. Bot. 12: 1-100. King, P. B. 1959. The evolution of North America. Princeton University Press, Princeton, New Jer- sey. Lawrence, G. H. M. 1951. Taxonomy of vascular plants. Macmillan & Co., New York. Meisner, C. F. 1841. Plantarum vascularium gen- era. Fasc. 10, tab. diagn. 313-344, comm. 225- 252. 1856. Polygonaceae. In: A. deCandolle, Pro- dromus systematis naturalist regni vegetabilis 14: 1-4, 28-186. Melchior, H. (ed.) 1964. Engler's syllabus der Pflanzenfamilien. 12th ed., Vol. II. Band An- giosperm. Gebriider Borntraeger, Berlin. Miers, J. 1851. On two genera of plants from the cordillera of Chile. Proc. Linn. Soc. London 2: 154-157. Moreau, R. L. P., and S. Crespo. 1969. Rehabilitacion de Eriogonum ameglinoi Speg. y caida en simonimia del genero Sanmartinia (Po- lygonaceae). Darwiniana 15: 112-114. Munz, P. A. 1974. A flora of southern California. University of California Press, Berkeley. Munz, P. A., and D. D. Keck. 1959. A California flora. University of California Press, Berkeley. Nuttall, T. 1817. Observations on the genus Eriogonum and the natural order Polygonaceae of Jussieu. J. Acad. Nat. Sci. Philadelphia 1: 24- 31, 33-37. 1848a. Descriptions of plants collected by Mr. William Gambel in the Rocky Mountains and upper California. Proc. Acad. Nat. Sci. Philadelphia 4: 7-26. 1848b. Description of plants collected by William Gambel, M.D., in the Rocky Mountains and upper California. J. Acad. Nat. Sci. Phila- delphia II, 1: 149-189. Parry, C. C. 1884. Chorizanthe R. Brown. Revision of the genus and rearrangement of the annual species, with one exception, all North Ameri- can. Proc. Davenport Acad. Nat. Sci. 4: 45-63. 1886. Harfordia Greene & Parry. A new genus of Eriogoneae from Lower California. Proc. Davenport Acad. Nat. Sci. 5: 26-28. Philippi, R. A. 1864. Plantarum novarum Chilen- sium. Linnaea 3: 225-228. 1873. Nuevas plantas en el herbario Chileno. Anales Univ. Chile 1873: 536. 1895. Plantas nuevas Chilenas de las familias que corresponden al tomo V de la obra de Gay. Polygoneas. Anales Univ. Chile 91: 487-526. Raven, P. H. 1963. Amphitropical relationships in the floras of North and South America. Quart. Rev. Biol. 38: 151-177. Remy, J. 1851. Botanica [Flora Chileana]. In: C. Gay, Historia fisica y politica de Chile 5(3): 257-284. Reveal, J. L. 1969a. A revision of the genus Eriogonum (Polygonaceae). Unpublished doctor- al dissertation, Brigham Young University Li- brary, Provo, Utah. 1969b. The subgeneric concept in Eriogonum (Polygonaceae), pp. 229-249. In: J. Gunckel (ed.), Current topics in plant science, Academic Press, New York. Reveal, J. L., and B. J. Ertter. 1976a. A reestab- lishment of the genus Stenogonum (Polyg- onaceae). Great Basin Nat. 36: 272-280. 1976b. Goodmania (Polygonaceae), a new genus from California. Brittonia 28: 427-429. Reveal, J. L., and J. T. Howell. 1976. Dedeckera 190 GREAT BASIN NATURALIST MEMOIRS No. 2 (Polygonaceae), a new genus from California. Brittonia 28: 245-251. Roberty, G., and S. Vautier. 1964. Les genres de Polygonacees. Boissiera 10: 7-128. Shreve, F., and I. L. Wiggins. 1964. Vegetation and flora of the Sonoran Desert. Stanford Uni- versity Press, Stanford, California. Small, J. K. 1898. Studies in North American Po- lygonaceae. I. Bull. Torr. Bot. Club 25: 40-53. Specazzini, C. 1902. Nova addenda ad Floram Patagonicam, 3. Anales Mus. Nac. Hist. Nat. Buenos Aires 7: 135-308. Stafleu, F.' A., et al. (eds.). 1972. International code of botanical nomenclature. Regnum Veg. 82: 1-426. Stebbins, G. L. 1942. Polyploid complexes in rela- tion to ecology and the history of floras. Amer. Naturalist 76: 36-45. 1974. Flowering plants. Evolution above the species level. Harvard University Press, Cam- bridge. Stebbins, G. L., and J. Major. 1965. Endemism and speciation in the California flora. Ecol. Monogr. 35: 1-35. Stokes, S. G. 1904. The genus Eriogonum. A study of its distribution and development in con- nection with the climatic changes in the West during and since the glacial period. Also a revi- sion of its species. Unpublished master's thesis, Stanford University Library, Stanford, Califor- nia. 1936. The genus Eriogonum. A preliminary study based on geographic distribution. J. H. Neblett Pressroom, San Francisco. Stokes, S. C, and G. L. Stebbins. 1955. Chromosome numbers in the genus Eriogonum. Leafl. W. Bot. 7: 228-233. Takhtajan, A. 1959. Die Evolution der Angiosper- men. Fischer, Jena. 1969. Flowering plants: Origin and dispersal. Oliver and Boyd, Edinburgh. Thorne, R. F. 1968. Synopsis of a putatively phy- logenetic classification of the flowering plants. Aliso 6: 57-66. TlDWELL, W. D., S. R. Rushforth, and D. Simper. 1972. Evolution of floras in the In- termountain Region, pp. 19-39. In: A Cronquist, et al., Intermountain flora: vascular plants of the Intermountain West. Vol. 1. Hafner Publ. Co., New York. Torrey, J. 1857. Description of the general bot- anical collections. In: A. W. Whipple, Reports of explorations and surveys to ascertain the most practicable and economical route for a railroad from the Mississippi River to the Pacif- ic Ocean. Vol. 4, part 4. A. O. P. Nicholson, Printer, Washington, D.C. 1858. Description of plants collected along the route by W. P. Blake. In: R. S. Williamson, Reports of explorations and surveys to ascertain the most practicable and economical route for a railroad from the Mississippi River to the Pacif- ic Ocean. Vol. 5, part 2. A. O. P. Nicholson, Printer, Washington, D.C. Torrey, J., and A. Gray. 1870. A revision of the Eriogoneae. Proc. Amer. Acad. Arts 8: 145-200. Watson, S. 1877. Descriptions of new species of plants with revision of certain genera. Proc. Amer. Acad. Arts 12: 246-278. 1879. Contribution to American botany, 2: Description of some new species of North American plants. Proc. Amer. Acad. Arts 14: 288-303. Welsh, S. L. 1974. Anderson's flora of Alaska and adjacent parts of Canada. Brigham Young Uni- versity Press, Provo, Utah. PROBLEMS IN PLANT ENDEMISM ON THE COLORADO PLATEAU Stanley L. Welsh Abstract.— The problem of distribution of plant endemics is explored, especially as regards distribution of these plant species on elevational and substrate bases. Endemics per unit area are greater at elevations above 1980 m and on finely textured soils. A summary of data and a list of references is presented. The Colorado Plateau has long been known as a center for endemic plants. The basis for that endemism most certainly in- volves the unique features of the Plateau Province, especially as regards climate, geo- logical history, and position along migratory routes. Each of these features has played a role, and the interpretation of the nature of endemism in the region must reflect the part each has played. Climate is considered as arid for the region generally, but the mountains within and bordering the Prov- ince stand as moist discontinuities in this arid middle portion of the Colorado River drainage system. Geology is, in large part, displayed in technicolor, with Mesozoic and Cenozoic strata dominating, and with the thin cover of vegetation hardly obscuring the geology. Only at higher elevations is the mantle of vegetation sufficiently dense as to cloak the geological substrate. Volcanic in- trusives and extrusives are limited in area. Sedimentary formations constitute the main features of the plateau region. Canyons of grand size are entrenched into the surfaces of the uplifted plateau, where formations are displayed along the margins of horizontal or gently dipping plateaus. Broad anticlines, synclines, or gentle to steeply plunging monoclines have been eroded to expose hogbacks or cuestas where vast stratigraphies are exposed in rel- atively short distances. Main migratory routes into the plateau involve the mountain sequences in the Rocky Mountains along both the eastern and western margins. The great river system has provided a lane for movement of prop- agules both up and down stream. Great Plains influence appears to have been mainly from the southwest. Hence, the flora of the plateau is a function of the sources of its plants; Mohavean, Chihuahuan, Great Plains, Rocky Mountain, Great Basin, etc. My own interest in plant endemics within the Colorado Plateau began more than two decades ago when I began research leading to a thesis on the vegetation of the Utah portion of Dinosaur National Monument (Welsh 1957). The geological control of vegetation at lower elevations in the plateau is readily apparent, but nowhere so overwhelming as along the margins of the Split Mountain anticline at Dinosaur Na- tional Monument. The interaction of various substrates to low annual precipitation through long periods of time has led to the demonstration of edaphic differences on a grand scale. Soil formation in a traditional sense is unknown. The substrate surface is often merely residual parent material only slightly modified from that a few inches be- low the surface. Alluvium and colluvium function differentlv from residual or parent materials. Conditions for growth on the substrates 'Life Science Museum and Department of Botany and Range Science, Brigham Young University, Provo, Utah 84602 191 192 GREAT BASIN NATURALIST MEMOIRS No. 2 available are often rigorous at best. Fre- quently, those plants capable of estab- lishment and reproduction are few in num- ber. Competition is therefore limited, with the harsh environments supporting very few higher plant species. This point is illustrated by a statement as- cribed to the venerable Walter P. Cottam concerning vegetation of the Mancos Shale formation. While camped on that formation, he is reported to have stated that, "there were only four species of plants growing here, and I don't know three-fourths of them.'' This statement, whether true or not, is indicative of the paucity of vegetation on some of the formation, and of the peculiar nature of some of the entities growing there. Worst possible substrates available are the clays derived from saline shales such as members of the Mancos Formation and its counterparts such as the Tropic Shale (Tu- nunk equivalent). Salt contents of 30,000 ppm are not uncommon. Additionally, the clays perform in a special manner when water is added. A rainstorm of some six hours' duration supplied from 2 to 6 inches of water in June 1972 to the Tropic Shale east of Glen Canyon City in Kane County, Utah. Penetration into the clay of the Trop- ic Shale did not exceed 2 inches. Even when wet to field capacity the clays pro- vide very little moisture for plant growth. Still, a few plants do grow on this substrate, some of them continuing to flower and to mature fruit at soil moisture levels below the arbitrary 15 atmosphere level accepted as the permanent wilting point. At higher elevations, where rainfall is greater, even these substrate types are over- whelmed by vegetation and a soil mantle is developed which eventually is effective in insulating the vegetative cover from the ef- fects of high salinity and from such sub- stances as selenium. Sandstones and their derivatives form the other extreme in edaphic situations. Soluble salts are fewer, selenium is more restricted, water penetration is greater, and the pro- portion of water which can be extracted from soil at field capacity is greater. In- creased species diversity on sands is appar- ently a function of the greater ease of es- tablishment and better all-around water and soil relationships, given equal elevation, temperature, humidity, and opportunity for establishment. Glaciation has apparently not played a major role in the plateau, except in the highlands along its fringes. Persistence of plants through the Pleistocene has been a possibility in at least the lower elevation portions, and indeed in most of the Plateau Province. With this background, it seems both fea- sible and necessary to attempt to quantify the data on endemism. The data have been generated from information recorded on herbarium labels of specimens deposited at Brigham Young University and at the Na- tional Museum of Natural History, Smithso- nian Institution. Further data have been taken from various taxonomic treatments (see list of references). Despite all attempts, the information available is fragmentary, and the conclusions derived from that infor- mation must be regarded as tentative. A portion of the problem involves the nature of endemics generally. Some of them are highly restricted and they have been col- lected only once or a very few times. Infor- mation on substrate, elevation, geological formation, and other pertinent information is not available either on the herbarium ma- terials or in the literature. Further, most plant taxonomists (and others who collect plants) are not well prepared in identi- fication of substrate types. The clays and silts of some investigators will be the sands and gravels of others. Despite the short- comings of the information, there seem to be trends which can form the basis for fur- ther investigations. The data summarized from the sources investigated are presented in Table I. Only 34 families of vascular plants are known to have endemic species in the region. The to- tal number of endemic entities is some 340, 1978 INTERMOUNTA1N BIOGEOGRAPHY: A SYMPOSIUM 193 an average of about 10 per family. The en- demies are far from equally divided among the various families. The Boraginaeeae (20), Compositae (58), Leguminosae (86), Polyg- onaeeae (28), and Scrophulariaeeae (29) con- tain some 221 species or about 65 percent of all endemic taxa. The greatest number of taxa occurs in the Legume family. The 86 entities constitute some 25 percent of all of those in the region. The genus Astragalus, with 62 endemic taxa, is not only the largest single contrib- utor to the list of endemics, but contributes a greater number of taxa than any family beside the legumes. Most of those taxa oc- cur at low elevations, contributing to the Table 1. Summary of plant endemics on the Colorado Plateau Elevation Family Name Taxa No. C, S, M Substrate Sand, G Both NA 1980 m <6500ft 1980m >6500ft Both NA Apocynaceae Asclepiadaceae Boraginaceae 4 4 20 1 3 3 2 7 1 2 10 4 2 9 1 3 4 4 Cactaceae 11 2 1 2 6 5 6 Capparidaceae Caryophyllaceae Chenopodiaceae Compositae Cruciferae 1 1 6 58 11 1 2 9 6 2 29 3 1 1 4 2 1 16 1 4 25 7 1 12 4 2 7 14 Cyperaceae Elaeagnaceae Ephedraceae Euphorbiaceae Gramineae 5 1 1 1 2 1 3 1 1 2 2 2 1 1 3 2 1 Gentianaceae 2 1 1 1 1 Geraniaceae 1 1 1 Hydrophyllaceae Labiatae 12 2 6 3 1 2 2 9 1 1 1 2 Leguminosae Astragalus 62 18 35 9 42 16 4 others 24 5 17 1 16 4 3 1 Liliaceae 4 4 3 1 Najadaceae Nyctaginaceae Onagraceae Papaveraceae 1 2 9 1 3 1 2 6 1 2 6 1 1 2 1 Polemoniaceae 13 3 10 5 5 2 1 Polygonaceae Eriogonum Other 27 1 14 12 1 1 17 8 2 Portulacaceae 1 1 1 Prirnulaceae 1 1 1 Ranunculaceae 6 6 4 2 Rosaceae 1 1 1 Rubiaceae 2 2 1 1 Scrophulariaeeae 29 7 21 1 8 13 8 Selaginellaceae Umbelliferae 1 12 3 1 9 1 5 5 2 Total 340 84 187 40 29 182 88 41 29 Percentage 24.7 55.0 11.8 8.5 53.5 25.9 12.1 8.5 Abbreviations: C, clay; S, shale, M, mud; G, gravel; NA, not available. 194 GREAT BASIN NATURALIST MEMOIRS No. 2 idea that endemism is especially great at low elevations. Despite this apparent abun- dance of endemics at low elevations, only slightly more than half of the entities are known from below 1980 m (6500 feet) in elevation; a quarter of them occur above 1980 m (6500 feet), 12 percent overlap, and 8.5 percent are unknown. However, the to- tal land area above 1980 m in elevation represents only about 30 percent of the re- gion. When endemics are expressed in per unit area figures there is 1 endemic per 1170 km2 for the higher elevations and 1 per 1318 km2 below that elevation. Thus, the incidence of endemics generally is greater at the higher elevations. The area of the Colorado Plateau is roughly 340,000 km2 (133,000 sq mi), which places the density of endemics at about 1 per each 1000 km2. Almost half (55%) of the endemics occur on sand and gravel, and a quarter on clays, shales, or muds. Those that grow on both types constitute only 11.8 percent, and data is missing for some 8.5 percent. It is difficult to derive data on the pro- portions of sands and gravels to other sub- strate types. Certainly the sands and gravels cover more of the land surface than do the other substrate types. If one assumes that only a quarter of the included region con- sists of clays, shales, and muds, then the en- demics on clays occur at about 1 per 1000 km2. Those on sands and gravels are present at a density of only 1 per each 1350 km2. Initial impressions by collectors that en- demics are especially rich on the finely tex- tured soils are supported by the fabricated data. Certainly if the total area occupied by clays is less than 25 percent, then the en- demics per unit area might be expected to be very great indeed. Surprising is the greater density of endemics in the montane sections of the plateau. Distribution of en- demics is hardly at random, and any at- tempt to provide averages or taxa per unit area is likely to obscure the geographic controls which result in the development and placement of the taxa. In a previous paper (Welsh, Atwood, and Reveal 1975), the endangered, threatened, extinct, endemic, and rare or restricted vas- cular plants of Utah are plotted by geo- graphic subdivision of the state. The large part of those plants represent endemic plants of Utah. Instructive from the plotting is the apparent unequal distribution of those plants in the state. Main centers of distribu- tion include the high plateaus of south cen- tral Utah, the Canyonlands section, and the Uinta Basin. A trend is present, however, which indicates that the greater numbers occur near the southern end of Utah, with progressively fewer taxa northward. Since endemic plants are likely to be considered as endangered or threatened, and since most of the lands within the Colorado Plateau are federally controlled, then some management of these specialties is in- dicated. Perhaps it is possible to establish some predictive guidelines with regard to location of endemics (i.e., where they might be expected to occur). Outcrops of shale, mudstone, and siltstone at all elevations should be considered as suspect sources for plant endemics. Any peculiar substrate, such as the lacustrine limestones at higher eleva- tions in the western margin of the plateau, should be considered as important localities for endemic taxa. Glaciated localities can generally be excluded from consideration in management practices. References Atuood, N. D. 1972. A revision of the Phacelia crenulata group (Hydrophyllaceae) for North America. Great Basin Nat. 35: 127-190. Barneby, R. C. 1964. Atlas of North American As- tragalus. Mem. New York Bot. Card. 13: 1-1188. Beaman, J. H. 1957. The systematica and evolution of Townsendia (Compositae). Contr. Gray Herb. 183: 1-151. Cronquist, A. J. 1947. Revision of the North Amer- ican species of Erigeron, north of Mexico. Brit- tonia 6: 121-302. Harrington, H. D. 1954. Manual of the plants of Colorado. Sage Books, Denver. 666 pp. Higgins, L. C. 1968. New species of perennial Cryptantha from Utah. Great Basin Nat. 28: 195-198. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 195 1971. A revision of Cryptantha subgenus Oreoearya. Brigham Young Univ. Sci. Bull., Biol. Ser. 13(4): 1-63. Kearney, T. H. and B. H. Peebles. 1951. Arizona Flora. University of California Press, Berkeley. McDougall, W. B. 1973. Seed plants of northern Arizona. Museum of Northern Arizona, Flag- staff. Tidestrom, I. 1925. Flora of Utah and Nevada. Contr. U. S. Natl. Herb. 25: 1-665. Welsh, S. L. 1957. An ecological survey of the veg- etation of the Dinosaur National Monument, Utah. Unpublished master's thesis, Brigham Young University, Provo, Utah. Welsh, S. L., N. D. Atwood, and J. L. Beveal. 1976. Endangered, threatened, ex- tinct, endemic, and rare or restricted Utah vas- cular plants. Great Basin Nat. 35: 327-376. SOME FACTORS GOVERNING PLANT DISTRIBUTIONS IN THE MOJAVE-INTERMOUNTAIN TRANSITION ZONE Susan E. Meyer' Abstract.— The existence of a floristic transition zone can be inferred by the fact that a high proportion of in- digenous plant species reach a distributional limit within the area. The vascular flora of Washington County, Utah, exhibits this character to a marked degree with 53.6 percent of the native flora reaching a distributional limit within the county. By looking at geographic distributions, ecological preferences, and range termination in- formation for the component species, first approximations are made as to the probable factors mediating plant distributions within the county, particularly of the range-terminating species. The high proportion of range-termi- nating species in the flora may be accounted for mainly by limiting factors associated with abrupt shifts along two environmental gradients. The first factor is climatic and is mediated largely by altitude; change occurs pri- marily along a north-south gradient. The second is both climatic and edaphic and is mediated by factors other than altitude per se; it is oriented in an essentially east-west direction. Species with narrower tolerances are shown to be more sensitive to these environmental shifts. Some of the species distributions are better explained by a model involving the effects of interactions between habitat mosaic and genetic homogeneity of given popu- lations on relative migration rates in the transition area. These species may have the capacity to migrate farther, but differences in migration rates give their distributional limits a quasi-stable aspect. These data suggest that species cannot simply be divided into those which are environmentally limited in their present distributions and those which are not. It seems more fruitful to regard these two conditions as extremes on a continuum which can be expressed as migration rate. A floristic transition zone is characterized by the range termination of a high propor- tion of the indigenous species. The vascular flora of Washington County, Utah, exhibits this character to a marked degree, with 53.6 percent of the 1,067 indigenous species reaching a distributional limit within the county. The objectives of this study are to obtain detailed documentation of plant dis- tributions within the transition area, to identify environmental factors that might be controlling range limits, and to elucidate relationships between the distribution pat- terns of groups of species and their ecologi- cal amplitude. Early observations on the floristic transi- tion zone as it occurs in Washington Coun- ty include those of Parry (1875), Merriam (1893), and Jones (1910). The nature of the zone as it occurs in Nye County, Nevada, has been examined in some detail by Beat- ley (1975), while Bradley (1967) has pub- lished a phytogeographic survey of Clark County, Nevada, which also lies along the zone. Additional Washington County obser- vations include those of Hardy (1947), Cot- tarn et al. (1959), and Woodbury (1933). Whittaker and Niering (1964) worked with problems of a similar nature in southern Arizona. Description of the Study Area Washington County is located in the ex- treme southwestern corner of Utah. It is a roughly rectangular area which spans an east-west distance of about sixty miles and a north-south distance of about forty miles. A 'Department of Biological Sciences, University of Nevada, Las Vegas. Present address: Department of Botany, Claremont Graduate School, Clare- mont, California 91711. 197 198 GREAT BASIN NATURALIST MEMOIRS No. 2 region of great environmental diversity, the conntv contains extensive outcrops of carbo- nate rocks, sandstones, shales, and both ex- trusive and intrusive igneous rocks (Cook 1960). Elevations range from about 2,000 feet (600 m) near the common boundary with Arizona and Nevada to over 10,000 feet (3100 m) at the top of the Pine Valley- Mountains to the north. The dramatic na- ture of the landscape is exemplified by Zion National Park, most of which lies within the county. These diverse environments are not min- gled randomly; they represent instead the junction and interdigitation of three larger land areas which are themselves relatively homogeneous. Figure 1 shows the geo- graphic position of the county relative to these areas and to the boundaries which separate them. The first of these boundaries separates the Basin and Range Province from the Colo- rado Plateau Province. These two areas are very different both lithologically and in the relative proportion of high to low land. The former is comprised of isolated mountain ranges composed mainly of carbonate rocks separated by broad, alluvium-filled valleys. The latter is comprised of a contiguous series of high plateaus composed mainly of sandstone and shale, with lower elevations restricted mainly to relatively narrow can- yon floors. The climate of the Colorado Plateau, as a whole, is more mesic than that of the Basin and Range Province. The tran- sition between these two physiographic- provinces takes place within the county. Superimposed on this boundary is another boundary, which is designated on Figure 1 as the 4,000-foot contour line. Valley floors to the north of this line consistently lie above 4,000 feet and thus have a climate which is comparatively cold. Valley floors to the south always always lie well below 4,000 feet and thus have a climate which is warmer. The line effectively divides the Ba- sin and Range Province into two subregions, the Great Basin region to the north and the desert lowland region to the south. The fact that part of Washington County lies south of this line indicates an attenuation of desert lowlands along the Virgin River drainage into an area of generally higher land. Methods The data presented here are summarized from an annotated checklist of the vascular plants of the study area (Meyer 1976). The systems of vegetational and floristic classifi- cation used here are more fully explained there. Each species was assigned to one of thirty-eight floristic groups on the basis of published distributional information. A flo- ristic group is considered as an assemblage of species which share a similar geographic- distribution, without historical or ecological implication per se. The floristic groups are here classified into two types of distribution patterns. An areal distribution classification aggregates the floristic groups into five classes on the basis of their areal extent. The directional distribution classification combines the flo- Fig. 1. Location of Washington County, Utah, relative to state and physiographic boundaries and to the 4000- foot contour line. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM ristic groups into seven classes on the basis of their main area of distribution relative to the study area. Introduced plants are not in- cluded in the tabulations. All species which are reported to reach a distributional limit within the county are designated as range-terminating species. The occurrence of each species in the nine major community types within the county was determined from field notes, herbarium specimens, and monographic- sources. It is assumed here that the occurrence of a species in a particular community type in- dicates something about its ecological am- plitude and more specifically about the cli- matic tolerance of the species. That is, the characteristic and the dominant plants of each community type are considered to be indicators of climate. It seems better to use community type as a climate indicator rather than elevation, because of the effect of local modifying factors on the relation- ship between elevation and climate. It would be even better to monitor environ- mental factors directly, but this was logisti- cally impossible. In order to measure the degree of habitat specialization exhibited by various segments of the flora, a fidelity index is used: total species in segment total community occurrences in segment FI =' total species total community occurrences The values in this equation are derived in the following way. Each species in the flora occurs in one to nine possible community types. The value of the denominator equals the total species number (1067) divided by the total number of community occurrences (2182), or 0.489, indicating that the average species in the flora occurs in approximately two community types (0.489 w xk). The value for the numerator is derived in similar fashion, using total species number divided by total community occurrence number for any subcategory or segment of the total flora. For example, the value of the numerator for the endemic class is 0.606 (83 species in segment divided by 137 total community occurrences). The value of FI is defined as unity for the total flora, since in that case numerator and denominator have the same value. The value of FI for the endemic class would be 1.24 (0.606 divided by 0.489). Since this val- ue is greater than unity, the endemic class shows a higher degree of habitat special- ization than the flora as a whole. Values smaller than unity values would indicate the converse. Given a value of 0.489 for the denomina- tor, the value of FI can range from 0.23 (0.111 or one-ninth divided by 0.489) to 2.05 (unity divided by 0.489). These two ex- tremes represent the respective cases of the average species in the segment occurring in all nine community types (one species/nine community occurrences) and in only one (one species/one community type). Species which appear to be limited to a single substrate are . designated as edaphi- cally restricted species. Halophytes are not included in the edaphically restricted cate- gory because halomorphic soils develop on a variety of substrates. Factors Governing M igration Rates Propagule dispersal is a stochastic pro- cess. A plant disperses propagules in all di- rections, regardless of whether the prop- agule is likely to be dispersed to a site favorable for its growth, or whether an ef- fective dispersal constitutes a range exten- sion for the species. An effective dispersal is accomplished when a propagule is dispersed to a site for which it is preadapted in terms of genetic- tolerance (Good 1930) and is therefore able to grow and reproduce. When this site falls outside the previous distributional area for the species, the process is called migration. Effective dispersal (and thus migration) is a function of three sets of variables. These 200 GREAT BASIN NATURALIST MEMOIRS No. 2 are distance, size of population supplying propagules, and time. Their operation is based simply on the laws of chance. The first two variables may be related to the third as an expression of rate. The rate at which a species is able to migrate is a func- tion of effective source-population size and of average minimum dispersal distance (the average minimum distance a propagule must be dispersed in order to arrive at a site for which it is preadapted). The larger the effective source-population size and the shorter the average minimum dispersal dis- tance, the higher the migration rate. Populations adjacent to the area of poten- tial colonization are more likely to disperse propagules into the area than those located at greater distances. Therefore, the rate of migration into an area is most influenced by properties of populations immediately adja- cent to that area. The size of effective source-population is conditioned by several factors: 1) absolute size and density of the population, 2) repro- ductive efficiency, and 3) dispersal effi- ciency. These simply say that the chance of propagules being dispersed to sites for which they are preadapted is increased if more propagules are produced and if prop- agules can disperse more efficiently. Minimum dispersal distance is the same for all propagules produced by a population only when they are all preadapted to iden- tical sites. More often, minimum dispersal distance is different for different propagules because of their preadaptation to different environments. The average minimum dis- persal distance represents a mean value for all the propagules produced. It is dependent upon three sets of factors: 1) environmental heterogeneity, 2) amount of genetic hetero- geneity for tolerance characters present in the population, and 3) degree of congruence between environment and tolerance charac- ters of the dispersed propagules. If the environment is homogeneous and favorable, the average minimum dispersal distance will be very short, and migration will be relatively rapid. If the environment is homogeneous and unfavorable, the aver- age minimum dispersal distance will be very long and migration will be relatively slow. Furthermore, these effects will in- crease with the degree of genetic homoge- neity of the population. Thus a population with narrow tolerance (low genetic heterogeneity) will be able to migrate rapidly through an environment which is homogeneous and favorable be- cause of the very high probability that a propagule will be dispersed to a site for which it is preadapted. But its rate of mi- gration through an environment which is homogeneous and unfavorable will approach zero, because of the very low probability that a propagule will be dispersed to a site for which it is preadapted. A population with broad tolerances (high genetic heterogeneity) would be able to mi- grate less rapidly than a species with nar- row tolerances through an environment which is homogeneous and favorable. This is because of the production of some prop- agules preadapted to conditions other than those present in the homogeneous environ- ment. It would, however, be able to mi- grate more rapidly through a homogeneous and unfavorable environment. This is be- cause it produces propagules preadapted to a wider variety of environments, and thus some propagules are likely to be preadapted for survival in the potential migration area. A heterogeneous environment is made up of a mosaic of different environments. A genetically heterogeneous population will produce propagules preadapted to a wide range of environments, and thus a high pro- portion of the heterogeneous environment will be "read" as favorable by the average propagule it produces. This means that the average minimum dispersal distance will be relatively small and thus that migration will be relatively rapid. A genetically homogeneous population will produce propagules preadapted to a narrow range of environments; hence a low proportion of the heterogeneous environ- ment will be read as favorable. Average 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 201 minimum dispersal distance will be relative- ly large, and migration will be relatively slow. This model is attractive in its simplicity but is lacking in several respects. Most im- portantly, it does not take into account the fact that both the genetic constitution of the population and the environment may themselves be changing through time. In ef- fect, it is an instantaneous rate expression; it simply describes the options at a given point in time. In nature the situation is complicated by the fact that several of the variables in the rate expression may them- selves be changing. A change in the environment might change the configuration of favorable and unfavorable sites for a population and thus alter its migration rate. In some cases sites already occupied might become unfavor- able, causing a diminution in range. Such secular climatic changes are well docu- mented in the general area under discussion (Mehringer 1965, Mehringer and Martin 1965). It is important to consider that a species may have attained its modern distri- bution under environmental conditions very different from those now prevailing. The genetic constitution of a population is subject to change from several sources. Two of these, gene flow and selection, are particularly important in this discussion. The former would tend to have the general effect of increasing heterogeneity. The lat- ter could have one of three effects. In a homogeneous and constant environment, normalizing selection would tend to de- crease variability. In a homogeneous but progressively changing environment, direc- tional selection would tend to produce an adaptive shift, providing that the original population was sufficiently variable. In a spatially heterogeneous environment, dis- ruptive selection could operate to produce genetically differentiated subpopulations, particularly if gene flow between sub- populations were minimized. These sources of genetic change are especially important in an area such as the one studied, which meets the criteria of Stebbins (1974) for an area in which active speciation may be ex- pected to be proceeding. It is also an area in which much of the habitat is hybrid- inducing in the sense of Anderson (1949). Many otherwise allopatric species have dis- tributions which overlap in this area; they can often be observed to introgress or inter- grade, depending on the viewpoint as to their origins. In spite of the shortcomings of the model of migration rate presented above, it is quite useful in interpreting Washington County data. Discussion of Data For interpretive purposes, two sorts of in- formation about each species are examined. One involves the distribution pattern for the species as a whole. The other involves eco- logical factors which govern its distribution within the transition area. When these two sets of information are examined together, a high correlation between geographic distri- bution and particular ecological require- ments becomes apparent. Table 1 shows the percent contribution of each of seven distributional classes to the range-terminating group as compared to the total flora. It can be seen that the classes do not contribute equally to these two groups. The Mojave, Colorado Plateau, and endemic classes contribute far more than their share to the range-terminating group, making up almost 70 percent of this group but com- prising only about 40 percent of the total flora. These differences are made up par- tially by the northern class, which is no- tably underrepresented in the range- terminating group. Column three of Table 1 shows the per- centage of species contained within each distributional class which are range-termi- nating. This value is almost 100 percent for the endemic and Mojave classes, indicating that a high proportion of the species may have reached a point which necessitates mi- gration through an environment that is 202 GREAT BASIN NATURALIST MEMOIRS No. 2 homogeneous and unfavorable. This possi- bility is reinforced by the fact that they show high fidelity indices, indicating a high degree of habitat specialization and thus a low degree of genetic heterogeneity for cli- matic tolerance characters. The Colorado Plateau and Great Basin classes also show a relatively high percent of range-terminating species, indicating a higher sensitivity to the environmental tran- sition between the Colorado Plateau and the Great Basin than might have been ex- pected. In fact, these two classes account for almost a third of all range-terminating species, a contribution almost as great as that of the Mojave class. The southern class also shows a high pro- portion of range-terminating species, but the converse is true of the northern group. These apparently do not read the environ- ment to the south of the county line as ho- mogeneously unfavorable. One explanation for this is the fact that the northern class is not as specialized (FI = < 1) and thus should be able to migrate more effectively through a heterogeneous or even homoge- neously unfavorable environment. Another possibility is that these species attained their distributions at some time in the past when environment was more homoge- neously favorable. The range-terminating members of the northern class have a much higher fidelity index than the class as a whole. These spe- cies would presumably have trouble mi- grating through the heterogeneous environ- ment of the transition zone because, due to their genetic homogeneity, they tend to read more of it as unfavorable. In fact, the range-terminating members of a class con- sistently show a higher fidelity index than the class as a whole, lending credence to the concept that low genetic variability for tolerance characters makes migration across a transition zone more difficult. Table 2 shows the percentage contribu- tion to each community type and to the to- tal flora by each of the seven directional distribution classes. The community types are arranged roughly in altitudinal se- quence. The nondirectional class is slightly to markedly overrepresented in every commu- nity type except the Hot Desert Shrub com- munity. In general, the species have broad tolerances and occur in a variety of commu- nity types, a fact also evidenced by their low fidelity index (0.79). About 85 percent of the Mojave species occur in the Hot Desert Shrub community, making up almost half of the total flora. Perhaps more significant is the fact that an appreciable percent of the species in two of the higher elevation communities, Foothill Woodland (11 percent) and Mountain Brush (6 percent) is comprised of Mojave species. Bearing in mind that over 98 percent of the Mojave species are range-terminating, it is clear that some of these species (almost 20 percent of the total Mojave class) reach a Table 1. Range termination and fidelity indices by directional class. % of Total % Range- Fidelity Index % Total Range- Terminating Fidelity Index Range- Directional Native Terminating Species Total Terminating Class Flora Species in Class Class Species Endemic Nondirectional Mojave Colorado Plateau (ireat Basin Northern Southern Anomalous Total Flora 7.8 25.5 17.6 14.3 6.6 20.5 5.5 2.2 100.0 14.0 <1.0 32.3 22.2 8.4 13.5 8.6 1.0 100.0 96.4 <1.0 98.4 83.0 68.6 35.2 83.1 53.6 1.24 0.79 1.32 0.86 1.26 0.96 1.04 1.24 1.34 1.12 1.46 1.22 1.09 1.23 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 203 termination of range in spite of the fact that their tolerance characters permit some individuals to disperse successfully to higher elevation community types. Most of the Mojave species have their main area of distribution in terms of num- bers of individuals in the Hot Desert Shrub community. Their occasional anomalous oc- currence in higher elevation community types indicates a broader climatic tolerance than that of species sharply restricted to the hot desert. Such species should be able to migrate through th^ ecotonal area because they can read at least some of it as favor- able. Therefore, there seems to be no a pri- ori reason for them to reach a limit of range in Washington County. Several fac- tors might operate, however, to make their rate of further migration very slow. Even though hot desert populations might be large, propagules preadapted for high elevation community types might be pro- duced at very low frequencies. In addition, populations occurring in higher elevation communities might have lowered reproduc- tive efficiency, thus retarding the estab- lishment of large populations which could produce a higher frequency of preadapted propagules. Both factors would have the ef- fect of keeping the effective source-popu- lation small. The migrating species might be able to occupy only a portion of the sites within the higher-elevation vegetation types; the proportion of favorable sites might decrease with increasing distance from the transition zone. This could be true if the species were only marginally tolerant climatically, if it had a narrow tolerance for some other envi- ronmental factor, or if it faced a com- petitive disadvantage in the presence of some high elevation species. All of these factors would have the effect of increasing the average minimum dispersal distance by increasing the proportion of unfavorable sites. Finally, the species may actually have migrated beyond the transition zone, but in such small numbers as to have escaped de- tection. In any case, the complex dynamics of the situation make it impossible to divide range-terminating species neatly into those which are limited by environmental factors and those which are not. The community contributions of the re- maining classes are easier to interpret. Since Rocky Mountain species enter the county via the Colorado Plateau, it is not surprising that the Colorado Plateau class is well rep- resented in montane communities, particu- larly those of middle elevations. Many of these montane species reach a western limit of range, probably because of the much lower proportion of favorable sites and the much larger average minimum dispersal dis- tances involved in migrating across the Great Basin. The class as a whole is under- represented in Hot Desert communities, and most species which do occur there are not range-terminating. The important role played by edaphic factors in this class will be discussed later. Very few of the Great Basin species enter the county at all; many reach southern and eastern range limits a few miles northwest Table 2. Percent composition of plant communities by directional class. % Total % % % % % % % % % Directional Native Hot Cold Desert Foothill Mountain Transition Mountain Mountain Mountain Class Flora Desert Desert Riparian Woodland Brush Forest Forest Riparian Meadow Endemic 7.8 6.6 7.0 2.9 10.1 8.2 7.8 4.7 2.7 4.6 Nondirectional 25.5 16.2 34.9 39.4 31.2 33.7 37.0 33.2 40.5 31.8 Mojave 17.6 46.0 9.5 17.3 10.7 5.8 1.0 1.0 1.4 0 Colorado Plateau 14.3 9.2 14.0 9.0 15.8 17.3 21.9 15.8 12.2 12.1 Great Basin 6.6 6.6 7.0 5.1 6.4 5.8 2.7 5.3 3.2 1.0 Northern 20.5 4.0 18.1 14.4 15.4 23.1 25.1 37.9 36.9 50.5 Southern 5.5 9.0 7.0 9.4 6.7 3.4 2.7 1.0 1.4 0 Anomalous 2.2 2.3 2.5 2.5 3.7 2.9 1.8 2.1 1.8 0 204 GREAT BASIN NATURALIST MEMOIRS No. 2 of the county line. Those which do enter the county show a high degree of habitat specialization, but, oddly enough, this spe- cialization is not to any particular commu- nity type. Classic Great Basin salt desert does not occur in Washington County, which may account for the dis- proportionately high representation of Great Basin species in other community types. Also, the total number of species is so small that a difference of a few species has a rela- tively large effect on the percent values. As would be expected, species of the northern class are abundantly well repre- sented in high montane communities, rea- sonably well represented in foothill commu- nities, and poorly represented in desert communities. Values for the southern class show the opposite relationship. They are poorly rep- resented in montane communities and well represented in desert communities. Members of the endemic class show a high degree of habitat specialization (FI = 1.24), and are overrepresented in relatively xeric communities at low and middle eleva- tions. They are noticeably underrepresented in riparian communities. This may be be- cause xeric environments tend to be more heterogeneous than mesic habitats in re- spect to variables other than moisture, and heterogeneous environments tend to restrict migration of specialized plant species, thus keeping them endemic. Possible relationships between the areal extent of a distribution type and the ecolog- ical amplitude of the component species will now be explored. Table 3 gives fidelity index values and percent community com- position for each of the five areal distribu- tion classes. These show that a distribution type which covers a large area is more likely to include species with broad toler- ances than a distribution type which covers a small area. The endemic and restricted classes show higher habitat specialization than the moderate class, which in turn shows a higher value than the wide class. This makes sense in light of the fact that species with broad tolerances should not only have more favorable sites available to them, but they should also be able to mi- grate more rapidly to occupy these sites. An exception to this trend is exhibited by the continental + class, which contains a high proportion of transoceanic species. The community composition values for species of this class show that they are heavily con- centrated in communities which provide aquatic habitat, and in fact many of these species are restricted to aquatic environ- ments. They have probably been able to achieve wide distributions in spite of being habitat specialists by virtue of superior dis- persal efficiency. It is these species which bring the index value for the Continental + group to a value which seems anomalously high. Table 3. Fidelity indices, percent edaphically restricted species, and percent composition of plant commu- nities by areal class. Endemic Restricted Moderate Wide Continental + % Total Native Flora 7.8 27.5 16.8 33.0 12.8 Fidelity Index 1.24 1.25 1.00 0.83 1.01 % Edaphically Restricted Species 36.1 39.8 15.7 7.2 0 % Hot Desert Shrub 6.6 44.5 23.4 21.4 1.7 % Cold Desert Shrub 7.0 21.3 17.8 43.5 7.9 % Desert Riparian 2.9 17.0 17.3 37.5 22.7 % Foothill Woodland 10.1 20.1 21.5 40.3 4.4 % Mountain Brush 8.2 14.9 21.2 46.6 6.3 % Transition Forest 7.8 18.3 15.1 46.1 11.0 % Mountain Forest 4.7 19.5 8.9 45.3 19.5 % Mountain Riparian 2.7 13.1 5.9 44.6 32.0 % Mountain Meadow 4.6 12.1 11.2 49.5 22.4 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 205 Plants of wide but subcontinental distri- butions do not show this specialization, but instead are well represented in all but the Hot Desert Shrub community type. Plants of moderate distribution tend to be better represented in the lower-elevation commu- nities. The trend continues in the group of restricted distribution, which is relatively underrepresented in all but the Hot Desert Shrub community type. This trend seems to indicate that groups of wider distribution show higher percentages in mesic types, while groups of narrower distribution show higher percentages in xeric types. This makes sense in view of the relative propor- tion of mesic to xeric environments in the world, on the continent, in the West, and in the intermountain-southwestern area, re- spectively. The values for the endemic class have been discussed above. The importance of edaphic factors in the transition area will be examined next. Toler- ance for edaphic factors, like tolerance for other environmental factors, is a matter of degree. This discussion takes into account only those species which have a very nar- row edaphic requirement. Even though these make up only about 8 percent of the total flora, the implication that edaphic fac- tors play a minor role in the transition area is not warranted. Many more species show some degree of edaphic specialization, and as noted above, distributional limits are not all or nothing propositions based on pres- Table 4. Percentage of all native species and of all edaphically restricted species occurring in each plant community. % Total % Edaphically Native Restricted Plant Community Species Species Hot Desert Shrub 32.4 63.9 Cold Desert Shrub 29.5 37.3 Desert Riparian 26.0 13.3 Foothill Woodland 27.9 32.5 Mountain Brush 19.5 15.7 Transition Forest 20.5 12.1 Mountain Forest 17.8 2.4 Mountain Riparian 20.8 2.4 Mountain Meadow 10.0 1.2 ence or absence of adaptive traits, but in- stead are also a matter of degree. The pat- terns for the sharply restricted species seem to indicate that edaphics may be more im- portant in the transition area than might be supposed. Table 4 shows the percentage of all spe- cies versus the percentage of edaphically re- stricted species occurring in each commu- nity type. Though both these values tend to decrease with an increase in altitude, edaphically restricted species drop much more sharply. In addition, this group is poorly represented in the Desert Riparian community. Clearly edaphically restricted species are much better represented in xeric than in mesic community types. This may indicate that the environment is more heter- ogeneous edaphically in dry than in wetter environments. Mesic soils developed on dif- ferent substrates tend to be more modified and more alike than xeric soils developed on those different substrates. A consequence of such environmental heterogeneity is that edaphically restricted species tend to belong to more restricted areal distribution types. This is because spe- cies with narrow edaphic tolerances migrate slowly through an edaphically hetero- geneous area. It may also be because the substrate itself is of restricted geographic occurrence. The first and third rows of Table 3 demonstrate this relationship. An edaphically restricted species will be able to migrate relatively rapidly in spite of low genetic heterogeneity as long as the en- vironment is homogeneous and favorable in terms of substrate. When such a species reaches a range limit, it is probably because it has reached a point where the substrate environment suddenly becomes very hetero- geneous or homogeneous and unfavorable. The favorable substrate may occur at wide- ly spaced sites of small areal extent instead of closely spaced sites of large areal extent, or it may be completely absent from the potential area. In any case, the average minimum dispersal distance becomes prohi- 206 GREAT BASIN NATURALIST MEMOIRS No. 2 bitively large, and migration slows to near zero. Table 5 shows the substrate preferences of edaphically restricted species belonging to each of the seven directional distribution classes. The northern class contributes no edaphicallv restricted species. This is not surprising in view of the high preference of this group for mesic environments, which do not promote edaphic specialization. The nondirectional class contributes only one limestone-restricted and one sandstone- restricted species, and as a group is marked- ly underrepresented. The southern class and the Great Basin class are also slightly under- represented and contribute only a handful of the edaphically restricted species. This means that most of these species must be contributed by the remaining three classes. Over a quarter of the species are contrib- uted by the Mojave class, but these are dis- tributed very unevenly among the sub- strates. Mojave species constitute almost two-thirds of the limestone-restricted class, about one-fourth of the sand-restricted class, and less than 10 percent to the remaining classes. The Colorado Plateau class accounts for almost a third of the total. But, in contrast to the Mojave class, none of these are lime- stone-restricted. Instead they make up al- most two-fifths of the sandstone-restricted species, one-fifth of the clay-restricted spe- cies, and about two-thirds of the sand- restricted species. Table 5. Edaphically restricted species by substrate a The endemic class accounts for the re- maining third of the total. This class con- tributes the only volcanic-restricted species, about three-quarters of the clay-restricted species, and two-thirds of the sandstone- restricted species. None of the endemics are sand restricted. Over 90 percent of the edaphically re- stricted species are range-terminating. The role of substrate specialization as a factor in range-termination seems clear. Limestone- restricted species appear to migrate rapidly across the southern Basin and Range Prov- ince, but their rate of migration nears zero when they encounter the sandstones and shales of the Colorado Plateau. Conversely, sandstone- and clay-restricted species seem to migrate rapidly across the Colorado Plateau, but their rate of migration nears zero when they encounter the high- carbonate soils of the Basin and Range Pro- vince. Sand is a substrate which is wide- spread, especially in the Colorado Plateau Province; thus sand-restricted species do not tend to be endemics. But there is sand in the Mojave Desert; thus some sand-restrict- ed species enter the county via the Mojave. These may be prevented from migrating further north by other than edaphic factors. Summary The data show that the high proportion of range-terminating species in the county flora is correlated with the abrupt shifts nd directional class. % Total % Edaphicallv % Volcanic- % Sand- % Sandstone- % Limestone- % Clay- Directional Native Restricted Restricted Restricted Restricted Restricted Restricted Class Flora Group Species Species Species Species Species Endemic Nondirectional Mojave Colorado Plateau Great Basin Northern Southern Anamalous Total Species % Total Edaphically Restricted Species 7.8 25.5 17.6 14.3 6.6 20.5 5.5 2.2 1()0.0 36.1 2.4 24.0 28.9 3.6 0 3.6 1.2 100.0 100.0 0 0 0 0 0 0 0 100.0 1.2 0 0 25.0 54.2 8.3 0 8.3 4.2 100.0 28.9 62.5 0 0 37.5 0 0 0 0 100.0 19.3 15.8 10.6 63.2 0 5.2 0 5.2 0 100.0 22.9 0 8.7 21.7 0 0 0 0 100.0 27.7 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 207 along environmental gradients oriented in north-south and in east-west directions. The north-south gradient is primarily climatic, while the east-west gradient is both climatic and edaphic. Species with narrower tolerances are shown to be more sensitive to these abrupt environmental shifts. It is also shown that some of the range-terminating species have the capacity to migrate beyond the transi- tion zone under present conditions, but that differences in relative migration rates give their distributional limits within the county a quasi-stable aspect. Conclusion Just about anything is statistically possible in biogeography. This makes it dangerous to conclude that any species has a migration rate of zero. It is possible for a propagule to be dispersed to a tiny enclave of favor- able environment in the midst of a vast area of homogeneous and unfavorable envi- ronment. And it is possible that the very rare propagule which has tolerance charac- ters near an extreme for the species will be dispersed to a site for which it is pre- adapted. These events may be very unlikely, but they are possible. They may result in some of the seemingly aberrant "dots" often seen on species distribution maps. Instead of considering only two possi- bilities, those species which are environ- ment-limited in their current distribution pattern and those which are not, it seems far better to consider these two possibilities as extremes on a" continuum. This involves recognition of the interplay among a com- plex series of factors which interact to pro- duce the probability that a species will be able to migrate at a given rate under a giv- en set of conditions. These concepts may prove difficult to tie down, but it is hoped the ideas presented here will facilitate the process— if they only begin to sharpen the focus on this complex and interesting prob- lem. Literature Cited Anderson, E. 1949. Introgressive hybridization. Hafner Pub. Co., New York. Beatley, J. C. 1975. Climates and vegetation pat- tern across the Mojave/Great Basin Desert tran- sition of southern Nevada. Amer. Midi. Natural- ist 93: 53-70. Bradley, W. G. 1967. A geographical analysis <>( the flora of Clark County, Nevada. Ariz. Acad. Sci. 4: 151-162. Cook, E. F. 1960. Geological atlas of Utah: Wash- ington County. Utah Geol. and Mineralog. Surv. Bull. 70: 1-124. Cottam, W. P., J. M. Tucker, and B. Drobnick. 1959. Some clues to Great Basin post-pluvial climates provided by oak distributions. Ecology 40: 361-377. Good, B. 1930. A theory of plant geography. New Phytologist 30(11):150-171. Hardy, B. 1947. The Beaverdam Mountains as a barrier in plant and animal distributions: addi- tional information. Proc. Utah Acad. Sci., Arts, and Lett. 24: 138 (Abstract). Jones, M. E. 1910. Origin and distribution of the flora of the Great Plateau. Contr. W. Bot. 13: 46-68. Martin, P. S., and P. J. Mehringer. 1965. Pleistocene pollen analysis and the biogeo- graphy of the Southwest, pp. 433-451. In: H. E. Wright, Jr. and D. G. Frey (eds.), The Qua- ternary of the United States. Princeton Univer- sity Press, Princeton, N.J. Mehringer, P. J. 1965. Late Pleistocene vegetation in the Mojave Desert of southern Nevada. J. Ariz. Sci. 3: 172-188. Merriam, C. H. 1893. The Death Valley expedition: a biological survey of parts of California, Ne- vada, Arizona, and Utah. N. Amer. Fauna 7: 1- 193. Meyer, S. E. 1976. Annotated checklist of the vas- cular plants of Washington County, Utah. Un- published master's thesis, University of Nevada, Las Vegas. Parry, C. C. 1875. Botanical observations in south- ern Utah in 1874. Amer. Naturalist 9: 14-21, 139-146, 199-205, 267-273. Stebbins, G. L. 1974. Flowering plants: evolution above the species level. Belknap Press, Cam- bridge, Mass. Whittaker, B. H., and W. A. Niering. 1964. Vegetation of the Santa Catalina Montains, Ari- zona. I. Ecological classification and distribution of species. J. Ariz. Acad. Sci. 3(1): 9-34. Woodbury, A. M. 1933. Biotic relationships of Zion Canyon, Utah, with special reference to succes- sion. Ecological Monogr. 3: 146-245. THE THEORY OF INSULAR BIOGEOGRAPHY AND THE DISTRIBUTION OF BOREAL BIRDS AND MAMMALS James H. Brown1 Abstract.— The present paper compares the distribution of boreal birds and mammals among the isolated mountain ranges of the Great Basin and relates those patterns to the developing theory of insular biogeography. The results indicate that the distribution of permanent resident bird species represents an approximate equilib- rium between contemporary rates of colonization and extinction. A shallow slope of the species-area curve (Z = 0.165), no significant reduction in numbers of species as a function of insular isolation (distance to nearest conti- nent), and a strong dependence of species diversity on habitat diversity all suggest that immigration rates of bo- real birds are sufficiently high to maintain populations on almost all islands where there are appropriate habitats. In contrast, the insular faunas of boreal mammals represent relictual populations that receive no significant con- temporary immigration. The insular mammal faunas have been derived by extinction from a set of species that colonized the islands when habitat bridges connected them to the continents in the late Pleistocene. A relatively steep species-area curve (Z = 0.326), no effect of isolation on species diversity, and the absence of appropriate species from large areas of apparently suitable habitat all support this conclusion. Measures of habitat diversity that are closely correlated with bird species diversity do not account for much of the variation in number of mammal species among islands. Insular area is the single variable that accounts for most of the variability in both bird and mammal species diversity; this supports the approach of using standard parameters such as area in comparative empirical analyses and general biogeographic theory. The results of this study suggest that extremes of vagility among taxa and a recent history of paleoclimatic and geological changes make it unlikely that equi- librial distributions, of the sort MacArthur and Wilson (1967) propose for the biotas of oceanic islands, are char- acteristic of the insular distributions of terrestial and freshwater vertebrates of western North America. Biogeography is an old science that has made recent advances by assimilating new concepts and data. Within the last two dec- ades information on continental drift, pa- leoclimatology, and ancient sea levels has revolutionized our understanding of histori- cal events and their effects on plant and an- imal distribution. During the same period, ecologists and evolutionary biologists have learned much about the processes of popu- lation growth, dispursal, extincton, speci- ation, and interspecific interactions which are the mechanisms that determine distribu- tion. Biogeography appears to be entering an exciting new period in which the volu- minous data acquired by systematists and descriptive biogeographers can be inter- preted in terms of recently understood his- torical events and ecological processes to draw quantitative relationships and derive general principles. With the publication of their equilibrium model of insular distribution, MacArthur and Wilson (1963, 1967) contributed not only a new theory but also a new approach to biogeography. Prior to their work, most biogeographic research had consisted of de- scribing the distributions of particular taxa and producing ad hoc, historical explana- tions. MacArthur and Wilson advocated a quantitative approach designed to build and test general models based on ecological pro- cesses. The specific model that they pro- posed suggests that the number of species inhabiting an island represents an equilib- rium between opposing rates of extinction and colonization, and that these processes are functions of the size of an island and its 'Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, Ariz 209 210 GREAT BASIN NATURALIST MEMOIRS No. 2 distance from a source of colonists, respec- tively. MacArthur and Wilsons's choice of is- lands for their revolutionary approach to biogeography was not a fortuitous one. Ever since the pioneering work of Darwin and Wallace, islands have played a preeminent role in the development of the science. It is difficult to do experiments in biogeography, but islands and insular habitats represent natural experiments; they are small, replica- ted systems among which species and envi- ronmental parameters are distributed in dif- ferent combinations. Often particular taxa are distributed among insular habitats in patterns that imply the operation of general mechanisms of dispersal, extinction, and in- terspecific interaction. The last decade has seen several attempts to test the model of MacArthur and Wilson, using a variety of organisms that inhabit both true islands and analogous isolated habitats (e.g., Barbour and Brown 1974, Brown 1971, Brown and Kodric-Brown in review, Culver 1970, Cul- ver et al. 1973, Diamond 1969, 1970a, 1971, Johnson 1975, Harper et al. this sym- posium, Schoener 1974, Seifert 1975, Sim- berloff 1974, Simberloff and Wilson 1970, Simpson 1974, Terborgh 1973, Vuilleumier 1970, 1973). Not all of these studies have supported the model, but the exceptions have contributed importantly to our under- standing of the patterns of insular distribu- tion and the historical events and ecological processes that produce them. Analysis of the distribution of vertebrates among isolated habitats in the Inter- mountain Begion of western North America has contributed significantly to the devel- opment and testing of biogeographic theory. The dedicated field work of several gener- ations of systematists has documented in de- tail the species distributions of most verte- brate groups (e.g., for fishes Miller 1948, Hubbs and Miller 1948, Hubbs et al. 1974, Smith, this symposium; for birds Behle 1943, 1955, 1958, this symposium, Grinnell and Miller 1944, Johnson 1965, 1970, 1973, 1974, 1975, Linsdale 1936, Miller 1935, 1946, Miller and Bussell 1956, Van Bos- sem 1936; for mammals Durrant 1952, Dur- rant et al. 1955, Grinnell 1933, Hall 1946). The paleoclimatic history of the region also is becoming increasingly well understood (e.g., Hubbs and Miller 1948, Martin and Mehringer 1965, Smith this symposium, Wells and Berger 1967, Wells and Jorgen- sen 1964). This excellent data base has been used in quantitative analyses of the insular distributions of lacustrine fishes (Barbour and Brown 1974) and montane birds (John- son 1975) and mammals (Brown 1971). Much more work remains to be done. Some of the most interesting contributions can be expected when additional kinds of organ- isms and insular habitats are studied, so that it is possible to make comparisons among taxa which differ in ecological requirements and dispersal abilities, and among habitats that differ in environmental parameters and history of isolation. The present paper discusses the distribu- tion of boreal birds and mammals among isolated mountain ranges in the Great Basin in relation to the theory of insular bio- geography. It attempts to relate distribu- tional patterns to mechanisms of dispersal and extinction and to differentiate relictual patterns that are the legacy of historical events from equilibrial ones that can be at- tributed to contemporary ecological pro- cesses. The paper tries to develop a general conceptual basis for analyzing and predict- ing insular distributions. It first discusses the current state of insular biogeographic theo- ry and then utilizes empirical data on the distribution of boreal birds and mammals to test the theory and search for general mech- The Theory of Insular Biogeography Colonization, extinction, and speciation are the primary processes that determine the composition of insular biotas. An island can be defined as a patch of suitable habitat surrounded by unfavorable environment that limits the dispersal of individuals. New spe- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 211 cies colonize an island either by dispersing across the habitat barriers or by immigrat- ing sometime when the barriers were tem- porarily absent. Speciation entirely within an island potentially is another source of new species, but there is no evidence that this process has contributed significantly to the diversity of boreal birds and mammals in the Great Basin region. Extinction, which may be caused by a variety of factors that reduce population size, eliminates species and reduces insular diversity. The composi- tion of an insular biota is determined by the interaction of these origination and extinc- tion processes. MacArthur and Wilson (1963, 1967) de- veloped a simple, general model of insular biogeography that represents the number of species inhabiting an island as an equilib- rium between contemporary rates of coloni- zation and extinction (Fig. 1). Their model makes the colonization rate a decreasing function of distance to a source of dis- persing species (usually the nearest conti- nent) and extinction rate a decreasing func- tion of island size. The model predicts: first, that relatively constant insular species di- versity is maintained by the continual turn- over (extinction and colonization) of individ- ual species; second, that the equilibrium number of species is positively correlated with island size and negatively correlated with distance to the source of potential col- onists; and third, that the equilibrium turn- over rate is inversely related to both island size and distance to a source of species. Al- though the MacArthur-Wilson model was designed specifically to account for the di- versity of organisms on oceanic islands, it has proven a useful heuristic device for ana- lyzing many kinds of insular distributions because it deals with general processes and makes robust, testable predictions. As the MacArthur-Wilson model has been tested using a variety of taxa distributed among both oceanic islands and several kinds of insular habitats (see references cited above), it has become increasingly clear that in its present form it is in- adequate to account for many empirical ex- \ COLONIZATION 1 X V 7 ¥ VEXTINCTIOINi .> A #^^ — \ """ NUMBER OF SPECIES °SF °LN NUMBER OF SPECIES Fig. 1. Two models of equilibrium insular biogeography. Left, the MacArthur-Wilson model, which portrays extinction and colonization rates as functions of island size and isolation respectively. Right, a modification of the MacArthur-Wilson model by Brown and Kodric-Brown (1977), which incorporates the effect of insular isolation on extinction rate. In both models intersections of the curves can be extrapolated to the abscissa and ordinate to give equilibrial numbers of species (S) and turnover rates (X), respectively. Note that the two models predict the same relative order of numbers of species but different orders of turnover rates with respect to island size and isolation. 212 GREAT BASIN NATURALIST MEMOIRS No. 2 amples. Two major problems have arisen in attempts to use the model to account for empirical patterns of animal and plant dis- tribution. First, in some insular habitats, his- torical episodes of immigration, speciation, and extinction have produced numbers of species that are significantly greater or less than the equilibrial number expected on the basis of contemporary rates of colonization and extinction (Barbour and Brown 1974, Brown 1971, Culver et al. 1973, Diamond 1970). Such historically determined distribu- tions should be particularly common in or- ganisms that are poor dispersers (such as the small, nonflying vertebrates) and in geo- graphic areas (such as western North Ameri- ca) where paleoclimatic and geological changes have drastically altered the barriers that currently isolate insular habitats. Sec- ond, it has proven difficult to test the Mac- Arthur-Wilson model's critical predictions about insular turnover. Although it is rela- tively easy to census accurately the biota of an island, it is much more difficult to assess the natural turnover rate. As a result, most of the purported measures of insular turn- overs have been criticized (see Diamond 1969, 1971, Terborgh 1973, Lynch and Johnson 1974, Simberloff 1974), and the critical predictions of how turnover rate varies with island size and isolation remain untested. Recently, Brown and Kodric- Brown (1977) obtained empirical evidence that insular extinction rates may be strongly dependent on the distance of an island from a source of colonists, rather than on island size alone as MacArthur and Wilson sug- gest. Areas that are sources of colonizing species often also may be sources of immi- grant individuals of species already present on the island; the arrival of these immi- grants may refuce the probability that in- sular populationd go extinct. We predicted that, for many islands, species turnover rates are lower for islands near sources of dis- persing species than for more isolated ones. Although this is the opposite of the pattern predicted by the MacArthur-Wilson model, the model can easily be modified to in- corporate this influence of immigration on extinction rate (Fig. 1). Neither of these problems detracts from the utility of MacArthur and Wilson's mod- el as a heuristic device or the value of their approach. Their work did much to stimu- late biogeographers to develop precise hy- potheses and test them with appropriately analyzed quantitative data. Gradually a con- ceptual understanding of insular bio- geography is emerging that may not be as simple and elegant as the MacArthur- Wilson model, but it is hoped it will be more realistic. In the following sections, I hope to use the distribution of boreal birds and mammals among isolated mountain ranges in the Great Basin to illustrate this theoretical approach and some of the result- ing concepts. Montane Islands The mountain ranges of the Great Basin are islands of coniferous forest and associ- ated mesic habitats in a sea of desert. The Great Basin is a vast interior drainage that lies between two montane continents, the central mountains of Utah (a part of the Rocky Mountains) on the east and the Sierra Nevada on the west (Fig. 2). Most of this area consists of broad arid valleys, which lie at an elevation of approximately 5000 feet and are sparsely covered with a vegetation dominated by low woody shrubs of the genera Artemisia, Chrijsothamntis, and Atriplex. Between the valleys are a series of mountain ranges oriented in a north-south direction. Many of these rise to over 10,000 feet; on their lower slopes they are covered with juniper-pinyon woodland and at higher elevations there are forests of mixed conifers, stands of aspens, and some- times wet meadows and permanent streams. The present analysis is based on 19 is- lands for which the boreal mammal fauna is adequately known; lists of boreal bird spe- cies are available for 13 of these. Islands were defined by operational criteria applied to topographic maps (U.S. Geological Sur- 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 213 vey maps of the states: scale 1:500,000). A montane island was considered to be a mountain range that contains at least one peak higher than 9800 feet and is separated from other highland areas by a valley at least 5 miles across below an elevation of 7500 feet. This altitude represents the ap- proximate lower boundary of juniper-pinyon woodland. For each island, area above 7500 feet elevation, distance to nearest continent (Sierra Nevada or central mountains of Utah), and elevation of highest peak were determined from topographic maps (Table 1). For 13 islands ornithologists (Johnson 1975, Behle, this symposium) have quan- tified the diversity of habitats available to boreal birds. This habitat diversity score in- corporates the number of coniferous tree species and the presence of riparian wood- land, wet meadow, and aquatic habitats (see Johnson 1975 for details). Since West (this symposium) has shown that juniper-pinyon woodland is absent or poorly developed in the northern Great Basin, the islands se- lected for this analysis lie south of the Humboldt River and Great Salt Lake to in- sure that they have somewhat comparable habitats. Boreal Birds and Mammals For the bird and mammal species that are restricted to juniper-pinyon woodland and more mesic habitats of higher eleva- tions, the mountains of the Great Basin are truly islands. Their boreal avian and mam- Table 1. Data for the boreal habitats used in the present analysis. Great Basin Montane "Islands" Area above 7,500 feet (sq. miles) Highest peak (feet) Nearest continent (miles) Habitat diversity score' 1 « 2 u & 1 .2 '3 u a -a 3 Small Boreal mammal species1 1. White-Inyo 738 14,242 10 11 8 11 2. Panamint 47 11,045 52 3 5 3 3. Desatoya 83 9,814 83 2 4 7 4. Toiyabe-Shoshone 684 11,788 110 7 6 13 5. Toquima-Monitor 1,178 11,949 114 - - 10 6. Roberts Creek 52 10,133 216 - - 4 7. Diamond 159 10,614 190 - - 4 8. Ruby 364 11,387 173 9 6 12 9. Spring 125 11,918 125 7 6 6 10. Sheep 54 9,912 86 5 5 3 11. Grant-Quinn Canyon 150 11,298 138 9 5 5 12. White Pine 262 11,188 150 - - 7 13. Schell Creek-Egan 1,020 11,883 114 - - 8 14. Spruce-South Pequop 49 10,262 156 4 4 4 15. Snake 417 13,063 89 14 9 10 16. Deep Creek 223 12,101 104 11 7 8 17. Pilot 12 10,704 114 - - 3 18. Stansbury 56 11,031 39 8 6 6 19. Oquirrh 82 10,626 19 9 6 6 Sierra Nevada Mainland 20. Carson 284 10,788 0 17 13 22 21. Yosemite 828 13,090 0 18 15 23 Rocky Mountain Mainland 22. Paunsaugunt-Aquarius 1,008 11,124 0 14 13 16 23. Uinta 1,536 13,498 0 15 14 21 'From Johnson (1975) and Behle (this symposium). 'See Appendix 1 for documentation. 'See Appendix 2 for documentation. 214 GREAT BASIN NATURALIST MEMOIRS No. 2 20 2 - BIRDS S= 2.526 A( r =0.701 20 50 100 200 500 1,000 20 U. 5 MAMMALS S = l 188 A0326 r =0.846 20 50 100 200 500 AREA ABOVE 7,500 FEET ELEVATION (SO MILES) 1,000 Fig. 2. Map of the Great Basin region of western North America showing the location of the isolated mountain ranges used in the present analysis. Numbers refer to individual montane islands and continental sample areas listed in Table 1. Note that islands of varying size and isolation lie in the "sea" of desert habitat between the central mountains of Utah on the east and the Sierra Nevada of California on the west. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 215 malian faunas are depauperate subsamples of the Sierra Nevada and Rocky Mountain faunas. All of the bird and mammal species that inhabit the isolated peaks are broadly distributed on one or (more frequently) both of the continental ranges to the east and west (see appendix and references cited there), but a significant proportion of the continental species are not present on any particular island and some are absent from all islands. The continental mountain ranges clearly are the source of the insular boreal faunas, and the distribution of species on the islands is the result of their abilities to colonize and avoid subsequent extinction. The definition of boreal bird and mam- mal species is somewhat arbitrary, and the lists of knowledgable specialists often do not agree precisely (e.g., compare the lists of boreal bird species of Johnson 1975, and Behle this symposium). I have followed Johnson's rather conservative designation of boreal bird species (for a list of these spe- cies and their distribution see Appendix 1). I have restricted my analysis to those spe- cies that he terms (again conservatively) "permanent resident Boreal species," be- cause I wanted to include in my com- parison with mammals only those species that maintain sedentary populations throughout the year. This eliminates from the analysis the large number of bird spe- cies that breed on the montane islands but are likely to migrate or disperse long dis- tances between successive breeding seasons. The definition of boreal mammal species in- cludes those that inhabit juniper-pinyon woodland or habitats of higher elevation, but not the desert habitats of the Great Ba- sin. I have excluded from the analysis large carnivores and ungulates because their dis- tributions have been drastically altered by human activity and their original ranges and habitat requirements are poorly known. Also I have ignored bats, because their dis- tributions are incompletely documented and they, like the migratory birds, probably are not permanent residents of the islands. The list of boreal mammal species (Appendix 2) is similar to that in Brown (1971), but it differs in some details because I have used slightly different criteria for designating bo- real species and I have included a few more records of occurrence. The resulting lists of boreal bird and mammals contain species of generally similar body sizes and habitat re- quirements, although there are approx- imately 50 percent more mammal than bird species on both the large islands and the continents. Species— Area Relationships The number of species (S) inhabiting an island usually is positively correlated with insular area (A); this relationship takes the form S = CAZ, where the values of the constant (C) and slope (Z) depend on the characteristics of the specific taxon and group of islands under consideration (Mac- Arthur and Wilson 1967, Preston 1962). The slope of this relationship can indicate the relative importance of extinction and origi- nation processes in determining the diver- sity of insular biotas. (Barbour and Brown 1974, Brown 1971, Mac-Arthur and Wilson 1967). Low slopes, Z<0.20, tend to charac- terize samples of different areas within a continent, and the Z-value varies with envi- ronmental heterogeneity (Harner and Har- per, 1977). Islands usually have higher Z- values; when the insular biota represents an approximate equilibrium between rates of colonization and extinction, Z-values tend to lie in the range 0.20-0.35. When there is no contemporary colonization to oppose extinc- tion, islands tend to have even higher Z-values, often > 0.40. The reason for this pattern is straightforward. On continents small sample areas contain a significant pro- portion of rare species. If these habitats were isolated and no immigration were per- mitted, rare species would go extinct rapid- ly on small islands and much more slowly on large ones, producing a much steeper species-area curve. Islands in equilibrium, where a significant rate of colonization op- poses extinction, represent an intermediate 216 GREAT BASIN NATURALIST MEMOIRS No. 2 situation; species that go extinct on small is- lands tend to be replaced by colonists, but not at a sufficient rate to produce diversity comparable to that on continents. There can be exceptions to this pattern. For ex- ample, if only a small subsample of the con- tinental biota is able to colonize a group of islands, the Z-value may be lower than ex- pected because large islands will not ac- quire as many species as they can support (Barbour and Brown 1974). This conceptual framework can be used to compare the species-area curves for the boreal birds and mammals. In my earlier analysis (Brown 1971), I argued that boreal mammals reached all of the islands during periods of climatic change in the late Pleistocene; since then there have been ex- tinctions but no colonizations. This con- clusion can be tested using a somewhat dif- ferent data set for mammals and comparing the mammalian and avian distributions. If birds are better dispersers and are currently crossing the desert valleys to colonize the isolated peaks, then they should have a sig- nificantly lower Z-value than mammals. This is the case. The number of species of both birds and mammals is correlated signif- icantly with area (Fig. 3, Tables 2 and 3). The Z-value for birds, 0.165, is even lower than that obtained for most insular biotas that are presumed to be in equilibrium and approximates values for continental samples. The Z-value for mammals, 0.326, is less than I obtained in my earlier analysis, Z = 0.428, and lies in the upper range of those observed for the biotas of true islands (Mac- Arthur and Wilson 1967). Thus species-area Fig. 3. The relationship of insular area to the number of permanent resident boreal bird species (above) and the number of small boreal mammal species (below). Note that the slope of the least squares regression line of the species-area curve for birds is much less steep than that for mammals. The equations for the fitted regres- sions and the correlation coefficients (r) are indicated. 1978 INTERMOUNTAIN HKK.l'ni.HM'HY: A SYMPOSIUM 217 curves for the boreal birds and mammals conform qualitatively to theoretical predic- tions, but they differ slightly from the quan- titative Z-values that might be expected if the avian distribution represents an equilib- rium between contemporary colonization and extinction, and the insular mammalian populations are Pleistocene relicts that do not disperse across the desert valleys. These slight deviations from expected Z- values are not difficult to explain. The low Z-value for birds suggests that the isolation of the montane islands may not be a signifi- cant barrier to avian colonization. The facts that a) approximately 80 percent of the var- iation in insular bird species diversity can be accounted for by the combined effects of area and elevation (Table 2) or by habitat diversity (Table 3), and b) there appears to be no impoverishment of species numbers resulting from isolation by distance to the nearest continent (Table 2, and see next sec- tion) suggest that colonization rates are high and there is little if any effect of insular isolation. This is consistent with Johnson's (1975) conclusion that the diversity of bo- real birds is attributable primarily to habi- tat; the impoverishment of the insular avi- faunas is the result of reduced habitat diversity on the isolated peaks and not to any significant extent to low colonization rates. It should be noted that it is not nec- essary to infer from this that the desert val- leys do not inhibit dispersal, only that colo- nization rates remain sufficiently high that boreal species which maintain breeding populations on the islands rarely go extinct and, if they do, they recolonize rapidly. The fact that the Z-value for mammals also is slightly lower than predicted has a different explanation. As I shall show in the next section, the mammals that have colo- Table 2. Stepwise multi birds and mammals inhabit iple in» regression of the influence of three variables montane islands in the Great Basin. on the number of species of boreal Variable1 Order entered in equation Birds Contribution to R2 F-value Significance level Area Highest peak Nearest continent 1 2 3 0.4915 0.2836 0.0051 10.633 12.607 0.201 0.008 0.005 0.658 Variable1 Order entered in equation Mammals Contribution to R2 F-value Significance level Area Highest peak Nearest continent 1 2 ^ 3 0.716 0.005 0.010 42.957 0.271 0.559 <0.001 0.610 0.466 'Data are log-transformed data from Table 1. Table 3. Correlation coefficients (r) between variables for the 13 montane islands for which all data are avail- able. Upper right matrix is log-transformed data; lower left matrix is computed with untransformed data. Note that number of bird and mammal species are not closely correlated with the same variables. Highest Nearest Habitat Bird Mammal Area peak continent diversity species species Area _ 0.795 -0.079 0.623 0.701 0.891 Highest peak 0.778 - -0.334 0.722 0.869 0.523 Nearest continent -0.083 -0.224 - -0.176 -0.367 -0.070 Habitat diversity 0.524 0.760 -0.102 - 0.851 0.539 Bird species 0.637 0.872 -0.310 0.898 - 0.611 Mammal species 0.876 0.817 0.090 0.542 0.592 - 218 GREAT BASIN NATURALIST MEMOIRS No. 2 nized the montane islands are those species of the continental fauna that occur at rela- tively low elevations within the boreal zone; species restricted to high elevations are absent from all of the isolated mountain ranges. The large islands, which also tend to have the highest peaks (Table 3), have ex- tensive areas of mixed coniferous forest, wet meadow, and other high altitude habitats. Even with post-Pleistocene extinctions these mountains would be expected to support some of the mammal species characteristic of these habitats on the continental ranges. Thus, if the islands had an unbiased sample of the boreal mammal fauna, the larger is- lands should have more species than they do and the species-area curve would be steeper. In my 1971 paper, I reported a higher Z-value than obtained here. The rea- son is that I included in the present analysis some marginally boreal species (Eutamias dorsalis, E. panamintinus, and Sylvilagus nuttalli) that are characteristic of juniper- pinyon woodland and are present on most of the montane islands. This had the effect of adding an approximately constant num- ber of species to the fauna of all islands, and thus lowering the slope of the exponen- tial species-area relationship. This is the same effect that is produced by the dis- proportionate representation of low eleva- tion species in the insular faunas relative to continental faunas because of differential colonization in the past. Isolation and Paleoclimatic History It is a common observation of insular biogeography that remote islands support fewer species than islands of comparable size and habitat diversity that are nearer to a continent (see MacArthur and Wilson 1967 and included references). This pattern is attributed to the limited ability of organ- isms to disperse so that the rate of immigra- tion to an island declines with increasing isolation. In the present analysis, neither birds nor mammals demonstrate such a negative relationship between number of species and distance to the nearest conti- nent (Tables 2 and 3). For mammals, this result is consistent with my 1971 analysis and conclusion that there is virtually no contemporary immigration to the isolated mountains because the desert valleys con- stitute almost absolute barriers to dispersal. The islands were colonized in the Pleisto- cene when periodic climate changes re- sulted in shifts in the altitudinal limits of the boreal vegetation. Data from plant macrofossils in woodrat middens suggest that, as recently as 8,000 to 12,000 years ago, periods of cooler, wetter climate en- abled juniper-pinyon woodland to flourish at least 2,000 feet (600 m) below its present lower elevational limit (Wells and Berger 1967, Wells and Jorgensen 1964). This was sufficient to make juniper-pinyon and asso- ciated meadow and riparian habitats con- tiguous across virtually all of the Great Ba- sin, and to enable the boreal mammals characteristic of these habitats to colonize all of the isolated mountain ranges. With the return of hotter, drier conditions, these "habitat bridges" connecting the islands to the continental ranges were eliminated and the insular mammalian faunas have been de- rived from the widespread Pleistocene fauna by independent extinctions on each island. Several lines of evidence are consistent with this interpretation. 1) The relatively steep species-area curve (Fig. 3) suggests that extinction has played a major role in determining mammalian diversity. This is supported by the discovery of late Pleisto- cene fossils of at least one boreal species (Marmota flaviventri.s) from an island (Spring Range) where it no longer occurs (Wells and Jorgensen 1964). 2) The lack of any correlation between number of mammal species and distance to nearest mainland (Tables 2 and 3) or any other likely measure of insular isolation (Brown 1971) suggests that there is no contemporary immigration to the islands. 3) All of the species known from the islands are found in juniper-pinyon or other habitats of com- parable elevation. It is possible to account 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 219 for the presence of the entire insular fauna in terms of colonization across habitat bridges that were known to have existed in the late Pleistocene. 4) All of the species of boreal mammals that are restricted to mixed coniferous forest or habitats of higher eleva- tion on the continental ranges are absent from all of the montane islands, even though large areas of apparently suitable habitat are present on some of the larger is- lands. These species include Maries ameri- cana, Aplodontia rufa, Eutamias alpinus, E. toivnsendi, E. speciosus, Tamiasciurus hud- sonicus, T. douglasi, Glaucomys sabrimis, Phenacomys intermedins, Clethrionomys gapperi and Lepus americanus. Paleobota- nies evidence indicates that the habitats of these species were not connected across the Great Basin during the late Pleistocene. This is further evidence that small terrestri- al mammals usually are unable to cross hab- itat barriers only a few miles in extent and thus colonize isolated habitats only when bridges of appropriate habitat provide di- rect access (Brown 1971, 1973). The explanation for the lack of correla- tion between number of species and dis- tance to nearest continent in birds appears to be the opposite of that in mammals: birds are such good dispersers that they have colonized virtually all islands with suitable habitat regardless of their isolation. Both Johnson (1975) and Behle (this sym- posium) report a negative relationship be- tween their measure of insular isolation (cu- mulative width of desert barriers) and number of permanent resident bird species, but both authors included continental sample areas in their analyses. The apparent effect of isolation in their analyses can be attributed largely to the fact that a signifi- cantly greater number of species inhabit continental sites than occur on the islands; there is little or no effect of isolation by distance when only islands are considered (see Fig. 2 of Johnson 1975). This is con- sistent with my own analysis of Johnson's and Behle's data. There is an insignificant negative correlation between number of bird species and distance to nearest conti- nent (Table 3), and the distance variable does not contribute significantly to account- ing for number of bird species using mul- tiple regression analysis (Table 2). It re- mains to explain the difference in species diversity between the continents and islands. Some of those species present on the conti- nents but absent from most of the islands may be limited in their distributions by their sedentary natures or their aversion to crossing inhospitable desert terrain. How- ever, many of these species such as Lagopus leucurus, Dryocopus pileatus, and Perisoreus canadensis, have specialized habitat require- ments and low population densities. Their habitats are totally lacking from many of the islands, and, even where they are pres- ent, they often consist of small patches ob- viously inadequate to support sustained populations. Two other sources of evidence suggest that rates of contemporary immigra- tion by most boreal bird species are suffi- ciently high to keep the habitats present on the islands filled with an appropriate com- plement of species. First, such boreal bird species as Picoides tridactylus and Cyano- sitta stelleri, which are restricted to high elevation, well-developed coniferous forests, are present on the islands. Since the habi- tats of these species have not been con- nected by bridges to the continental ranges during the Pleistocene, it must be inferred that at least these species are able to colo- nize across significant barriers of unsuitable habitat. Second, vagrant individuals of some boreal species (e.g., Cyanositta stelleri and Cinclus rnexicanas) infrequently are report- ed significant distances from breeding popu- lations (Johnson 1975, B. Bundick, pers comm.). This evidence indicates that even relatively sedentary permanent resident bo- real birds are much more vagile than small boreal mammals. Determinants of Species Diversity and Composition Differences in dispersal account in large part for differences in the species-area rela- 220 GREAT BASIN NATURALIST MEMOIRS No. 2 tionships between birds and mammals and in faunal composition between insular and continental mountain ranges (at least for mammals), but the considerable variation in species diversity among the insular bird and mammal faunas must be attributed primari- ly to ecological characteristics of the is- lands. Although a common set of both bird and mammal species has had the opportu- nity to colonize virtually all islands, the iso- lated mountain ranges support different numbers and kinds of species. Species diver- sity of both birds and mammals is closely correlated with insular area, but area is only a correlate of factors such as the quan- tity and diversity of habitat and food re- sources that ultimately determine species di- versity. Power (1972) and Johnson (1975) have shown for birds that someone familiar with a group of organisms can devise quan- titative measures of habitat diversity that account for significantly more of the vari- ance in insular species diversity than area. This approach has great utility for elucidat- ing the environmental factors that influence the diversity and distribution of particular taxa, but it may not be useful for devel- oping general biogeographic or ecological theory. There are two arguments in favor of bas- ing biogeographic theory on simple, stan- dard parameters such as insular area. The first is practical. Parameters such as area, elevation (another correlate of habitat diver- sity), and distance to the nearest continent are easy to determine from maps. By ob- taining these measurements and species lists from the literature, it is possible to describe clear patterns of plant and animal distribu- tion without doing all the original fieldwork required to quantify accurately more direct environmental variables. Much of the recent synthetic work in insular biogeography has been done this way, and even those authors that have attempted to quantify variables such as habitat diversity often have used data available from the literature or topo- graphic maps (e.g., Power 1972, Johnson 1975). The second argument is that area may be the best parameter for constructing general theory that can be applied to di- verse taxa and various kinds of islands. In the development of theory some sacrifice of precise explanation of specific cases usually must be made in order to obtain a desirable degree of generality. It is questionable whether it would be practical or profitable to base biogeographic theory on organism- specific parameters such as habitat diversity, food availability, or carrying capacity. Cer- tainly at present we have no accepted, stan- dardized techniques for measuring these variables in the field that are suitable for a variety of taxa and habitats. Some test of the merit of this approach can be made by comparing the correlates of species diversity for boreal birds and mam- mals. Johnson (1975) developed a quan- titative "habitat diversity score" that ac- counted for most of the variability in the number of boreal bird species in his analy- sis. This appears to measure accurately the requirements of boreal birds in western North America, because Behle (this sym- posium) tested it and obtained gratifyingly similar results. Since boreal mammals utilize the same forest, meadow, and freshwater habitats as boreal birds, it would be encour- aging if Johnson's habitat diversity score ac- counted for similarly large proportions of the variability in insular species diversity for both taxa. Unfortunately this is not the case (Table 2). The number of boreal bird and mammal species inhabiting the same is- lands are not very closely correlated (r2 < 0.37); whereas 81 percent of the variability in bird species diversity can be attributed to habitat diversity score, the comparable fig- ure for mammals is only 29 percent. Area is by far the best correlate of species diversity for both birds and mammals. When the ap- propriate log-transformed data are used, area accounts for 49 and 79 percent of the variability in bird and mammal species numbers, respectively (Table 3). Inter- estingly, elevation of highest peak (another correlate of habitat diversity readily obtain- able from topographic maps) and area taken 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 221 together in multiple regression analysis (Table 2) account for almost as much varia- bility in bird species diversity (r2 = 0.76) as the habitat diversity score (r2 = 0.81). Two explanations can be offered for the lack of good correlation between Johnson's habitat diversity score and mammalian spe- cies diversity. First, the habitat diversity score is highly dependent on the species di- versity of coniferous trees. Particular tree species and forest types probably are impor- tant to the boreal birds, but they are much less likely to influence the diversity of bo- real mammals. As I have shown, boreal mammals that depend on the well-developed coniferous forests of high elevations have been unable to colonize any of the islands, and those mammals that have colonized, if they require conifers at all, are primarily species of the juniper-pinyon woodlands that are well developed on almost all of the islands. Second, Johnson's habitat diversity score is based primarily on the presence or absence of particular tree species or habitat types. This qualitative approach appears to work well for birds, because they have suffi- ciently high immigration rates to maintain populations on almost all islands where suit- able habitats are present. In contrast, mam- mals, which have had to maintain insular populations for thousands of years in the absence of any significant immigration, may be much more dependent on the quantity of boreal habitats (particularly those character- istic of low elevations that were connected in the Pleistocene) which may be more closely correlated with insular area than with Johnson's more qualitative index. Whether one or both of these nonexclusive explanations turn out to be correct, it is clear that different factors determine the di- versity of insular species of birds and mam- mals. These results suggest that parameters such as area, even though their effects are indirect and imprecise, may be the variables best suited for constructing and testing gen- eral biogeographic theory. However, specific factors that influence the insular species diversity and composition of particular taxa can contribute to theory by providing information on the mecha- nisms and effects of the underlying process- es of colonization and extinction. The great importance of habitat in determining insular bird species diversity, coupled with a low slope of the species-area curve and the lack of any significant effect of isolation by dis- tance, suggests that boreal birds are able to disperse at a sufficient rate to maintain populations on most islands with suitable habitats. The distribution of boreal mam- mals presents a dramatic contrast. Mammal species are absent from many islands that have suitable habitats, either because they have never colonized or because they have gone extinct since the last episode of Pleistocene colonization. The interacting ef- fects of habitat requirements and paleocli- matic changes in determining the set of spe- cies that colonized the islands already have been discussed. Of equal interest are the factors that have resulted in extinctions among the original colonists to produce the present insular faunas. The mammalian faunas of the montane is- lands have been derived by extinction from a common set of 14 functional species that were widely distributed across the Great Basin during the late Pleistocene when their habitats were connected by bridges to the continental mountain ranges (Table 4). After the islands were isolated by the contraction of boreal habitats to approximately their present position, extinctions reduced the faunas of each island and thus played a ma- jor role in determining species composition. Five large islands have retained at least 10 of their original 14 species, but five small islands have lost all but 3 or 4 of the origi- nal set of species in the period of approx- imately 10,000 years that the boreal habi- tats have been isolated. The distribution of extinctions among species is highly non- random and appears to be related primarily to population size, as MacArthur and Wil- son (1967) predicted. Herbivorous species of generalized habitat requirements and small to intermediate body size have persisted on 222 GREAT BASIN NATURALIST MEMOIRS No. 2 most of the islands (Table 4). In contrast, herbivores of large body size and/or spe- cialized habitat requirements and carnivores have had higher extinction rates and persist on only a small proportion of the 19 islands. The frequency of occurrence of boreal spe- cies on the islands (Table 4) corresponds very closely to the relative abundance of these species where they occur together on large islands or continental mountain ranges (personal observations). The dependence of extinction on population size probably is even more precise than is apparent here, because population size is influenced by the presence of particular habitats and com- peting species which vary among islands. For example, one of the few mountain ranges where a juniper-pinyon chipmunk (Eutamias dorsalis or E. panamintinus) does not occur is the Ruby Mountains, where juniper-pinyon woodland is very poorly de- veloped. Similarly, the chipmunk character- istic of higher elevations (Eutamias um- brinus) is absent from the Pilot Range, which has only small stands of mixed con- ifers on its single tall peak. On this moun- tain, in the absence of E. umbrinus, E. dor- salis has extended its altitudinal range upward into the mixed conifers on the peak. Conclusions, Predictions, and Unanswered Questions The insular distributions of boreal birds and mammals in the Great Basin differ in ways that are consistent with both current biogeographic theory and independent evi- dence of ecological and historical factors that have affected colonization and extinc- tion. All data seem consistent with the in- terpretation that insular bird populations represent an equilibrium between contem- Table 4. Characteristics of the boreal mammal species that inhabit the montane islands of the Great Basin. Species are listed in decreasing order of frequency of occurrence. Body Number weight of islands Species (grams) Diet Habitat inhabited Eutamias umbrinus 60 mostly seeds generalist: forests, wood- lands, talus 17 Neotoma cinerea 300 green vegetation generalist: rock outcrops or talus in all habitats 17 Eutamias dorsalis, E. 55 mostly seeds generalist: primarily 16 panamintinus juniper-pinyon woodlands Spermophilus lateralis 170 green vegetation seeds generalist: open forest, meadow 14 Microtus longicaudus 45 green vegetation generalist: meadow, open forest, streams 13 Sylvilagus nuttallii 800 green vegetation generalist: all habitats except dense forest 12 Marmota fhviventris 3,000 green vegetation generalist: open forest, meadow 10 Sorex vagrans, S. 7 invertebrates generalist: forest, meadow. 8 tenellus' streams Sorex palustris 14 invertebrates permanent streams 6 Oehotona princeps 120 green vegetation talus adjacent to meadow 5 Zapus princeps 25 seeds, green vegetation wet meadow, streams 4 Mustela erminea .50 small vertebrates generalist: meadow, open forest 4 Spermophilus beldingi 300 green vegetation wet meadows 3 Lepus townsendii 3,000 green vegetation large open meadows 1 Congeners listed on the same line are ecological and geographic replacements. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 223 porary rates of colonization and extinction; immigration rates are sufficiently high that boreal bird populations inhabit almost all is- lands where there are suitable habitats, re- gardless of their isolation from continents. In contrast, the boreal mammal populations represent relicts of species that were wide- spread in the Great Basin during the late Pleistocene when habitat bridges connected the islands and continents; the mammal faunas of the islands have been derived from a common set of Pleistocene colonists by subsequent extinctions in the absence of immigration. Neither the avian nor the mammalian dis- tributions fit the equilibrium model pro- posed by Mac Arthur and Wilson (1967) to account for diversity on oceanic islands. The insular mammal faunas clearly are not in equilibrium; in the absence of immigra- tion they are gradually relaxing toward an equilibrium of zero species at rates in- versely related to island size. The distribu- tion of birds does represent a sort of equi- librium between contemporary colonization and extinction, but the immigration rates are so high that the islands are virtually sat- urated with species for which the appropri- ate habitats are present. The avifaunas of the islands differ from the MacArthur-Wil- son model in that there is no significant ef- fect of isolation by distance, species compo- sition is quite precisely determined by habitat, and rates of faunal turnover prob- ably are very low (see Brown and Kodric- Brown 1977). The information available on the distribu- tion of other vertebrates among insular hab- itats suggests that the kinds of patterns de- scribed here for birds and mammals may be widespread. Migratory birds, bats, and very large carnivorous and herbivorous mammals probably are at least as vagile as permanent resident boreal birds. They would be ex- pected to colonize most suitable habitats and show little effect of insular extinction or isolation by distance. On the other hand, amphibians and reptiles probably have dis- persal capacities similar to mammals. These taxa are unlikely to cross habitat barriers of even modest extent, and it is well known that fishes can colonize new areas only when suitable habitat bridges are present (Barbour and Brown 1974, Hubbs and Mill- er 1948, Hubbs et al. 1974, Smith, this sym- posium). The distributions of these taxa should be extremely sensitive to paleocli- matic and geological events; where insular habitats have been connected and reisolated they should show relictual distributions in which the identity of the colonists and the effects of subsequent extinctions produce patterns of diversity comparable to those of the small boreal mammals. It is interesting to speculate that the extremes of vagility which characterize most of the vertebrates and the extensive climatic and geological changes that have drastically changed the landscape of western North America within the last million years make it unlikely that land and fresh water vertebrates in this re- gion will show the sort of equilibrium distri- butions that have been demonstrated for the biotas of oceanic islands with a long history of isolation and relative environmental sta- bility (see MacArthur and Wilson 1967, Simberloff 1974). In insular biogeography many important questions remain to be answered, and much theoretical and empirical work is yet to be done before this promising and vigorous young science can afford to rest on its lau- rels. The vertebrates that are distributed among the numerous isolated habitats in western North America continue to offer great potential as systems for testing theory and developing general concepts. For those organisms that appear to demonstrate equi- librial distributions it is particularly impor- tant to measure turnover rates accurately, and to incorporate the results into an ap- propriate conceptual framework. For those organisms that exhibit relictual distributions it is important to elucidate patterns of colo- nization and extinction and to relate those to theory. Three generations of icthyologists (Hubbs and Miller 1948, Hubbs et al. 1974, Smith, this symposium) have carefully dis- 224 GREAT BASIN NATURALIST MEMOIRS No. 2 sected historical changes in Great Basin drainages and have related them to the con- temporary distribution of fishes. As yet there has been little attempt to relate these patterns to biogeographic theory, but the possibilities for doing so seem great and it is hoped it will be attempted soon. The distri- butions of many organisms are patchy on scales smaller than the gross biogeographic one considered here. Smith (1974a, b) has obtained some encouraging results by using the theory of insular biogeography and the dynamics of opposing colonization and ex- tinction rates to account for the distribution of a small boreal mammal (Ochotona prin- ceps) among the isolated patches of its spe- cialized rockslide habitat in the Sierra Ne- vada. It will be interesting to see to what extent this approach continues to prove use- ful for understanding insular distributions on various scales. Acknowledgments This paper could not have been written without the field work and published spe- cies lists of numerous collectors, system- atists, and biogeographers who have worked to document the distributions of birds and mammals in the Great Basin. I am particu- larly grateful to N. K. Johnson for providing most of the data on birds and to W. H. Behle for making available information presented in this symposium. My wife, A. Kodric-Brown, and my students have done much to encourage my work in insular biogeography. I would like to dedicate this paper to the memory of Stephen D. Dur- rant, whose knowledge of the Great Basin and its mammals was matched only by his enthusiasm for the outdoor life and his abil- ity as a teacher and storyteller. Appendix 1. Records of occurrence ot permanent resident boreal bird species.' Sample Area Species 1 2 3 4 8 9 10 11 14 15 16 18 19 20 21 22 23 Dendragapus obscurus Bonasa umbellus Lagopus leucurus Oreortyz picta Glaucidium gnotna Strix occidentals S. ncltulosa Drijocopus pileatus Dendrocopus villostis D. pubescens D. albolaruatus Picoides arcticus P. tridactylus Perisorcus canadensis Cyanositta stellch Varus gambeli Sitta carolinensis S. pi/gmaea Cinclus mexicanus Pinicola cnucleator Total 5 4 6 6 6 5 6 13 15 'Data from Johnson (1975) and Behle (this symposium). 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 225 » 2 = 5 »-~ = ;£ j: •§ E § | J =~' ~ ~ 8. £ = -- ^ 2 > a. c £ \%\ llllli | 111 ill 111! lillj ill fill iff 6 | | ^11 226 GREAT BASIN NATURALIST MEMOIRS No. 2 Literature Cited H. Brown. 1974. Fish spe- lakes. Amer. Naturalist 108: Barbour, C. D., and J. cies diversity in 473-489. Behle, W. H. 1943. Birds of Pine Valley Mountain Range, southwestern Utah. Univ. Utah Biol. Ser. 7(5): 1-85. 1955. The birds of the Deep Creek Moun- tains of central western Utah. Univ. Utah Biol. Ser. 11(4): 1-34. 1958. The birds of the Raft River Mountains, northwestern Utah. Univ. Utah Biol. Ser. 11(6): 1-40. 1978. Avian biogeography of the Great Basin and Intermountain Region. Great Basin Nat. Mem. 2: 55-80. Brown, J. H. 1971. Mammals on mountaintops: nonequilibrium insular biogeography. Amer. Naturalist 105: 467-478. 1973. Species diversity of seed-eating desert rodents in sand dune habitats. Ecology 54: 775- 787. Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction.' Ecology 58: 445-449. Culver, D. C. 1970. Analysis of simple cave com- munities. I. Caves as islands. Evolution 29: 463- 474. Culver, D. J. R. Holsinger, and R. Baroody. 1973. Toward a predictive cave biogeography: the Greenbrier Valley as a case study. Evolution 2: 689-695. Diamond, J. M. 1969. Avifaunal equilibria and spe- cies turnover rates on the Channel Islands of California. Proc. Nat. Acad. Sci. 64: 57-63. 1970. Ecological consequences of island colo- nization by Southwest Pacific birds. II. The ef- fect of species diversity on total population den- sity. Proc. Nat. Acad. Sci. 67: 1715-1721. 1971. Comparison of faunal equilibrium turnover rates on a tropical and a temperate is- land. Proc. Nat. Acad. Sci. 68: 2742-2745. Durrant, S. D. 1952. Mammals of Utah, taxonomy and distribution. Univ. Kansas Pnbl. Mus. Natu- ral Hist. 6: 1-549. Durrant, S. D., M. R. Lee, and R. M. Hansen. 1955. Additional records and extensions of known ranges of mammals from Utah. Univ. Kansas Publ. Mus. Natural Hist. 9: 69-80. Grinnell, J. 1933. Review of the recent mammal fauna of California. Univ. California Publ. Zool. 40: 71-234. Grinnell, J., and A. H. Miller. 1944. The distri- bution of the birds of California. Pac. Coast. Avif. 27: 1-608. Grinnell, J., and T. I. Storer. 1929. Animal life in the Yosemite. Univ. of California Press, Berke- ley. Hall, E. R. 1946. Mammals of Nevada. Univ. Cali- fornia Press, Berkeley. Harner, R. F., and K. T. Harper. 1977. The role of area, heterogeneity, and favorably in plant species diversity of pinyon-juniper ecosystems. Ecology 57: 1254-1263. Harper, K. T., D. C. Freeman, W. K. Ostler, and L. G. Klikoff. 1978. The flora of Great Basin mountain ranges: diversity, sources, and dis- persal ecology. Great Basin Nat. Mem. 2: 81- 103. Hubbs, C. L., and R. R. Miller. 1948. Correlation between fish distribution and hydrographic his- tory in the desert basins of western United States, p. 17-166. In: The Great Basin with em- phasis on glacial and postglacial times. Bull. Univ. Utah 38(2): 1-191. Hubbs, C. L., R. R. Miller, and L. C. Hubbs. 1974. Hydrographic history and relics fishes of the north central Great Basin. Memoirs Calif. Acad. Sci. 7: 1-259. Johnson, N. K. 1965. The breeding avifaunas of the Sheep and Spring ranges in southern Nevada. Condor 67: 93-124. 1970. The affinities of the Boreal avifauna of the Warner Mountains, California. Occas. Pap. Biol. Soc. Nevada 22: 1-11. 1973. The distribution of Boreal avifaunas in southern Nevada. Occas. Pap. Biol. Soc. Nevada 36: 1-14. 1974. Montane avifaunas of southeastern Ne- vada: historical change in species distribution. Condor 76: 334-337. 1975. Controls of number of bird species on montane islands in the Great Basin. Evolution 29: 545-657. Linsdale, J. M. 1936. The birds of Nevada. Pac. Coast Avif. 23: 1-145. Lynch, J. F., and N. K. Johnson. 1974. Turnover and equilibria in insular avifaunas, with particu- lar reference to the California Channel Islands. Condor 76: 370-384. MacArthur, R. H., and E. O. Wilson. 1963. An equilibrium theory of insular biogeography. Ev- olution 17: 373-387. 1967. The theory of island biogeography. Princeton Univ. Press, Princeton, N. J. Martin, P. S., and P. J. Mehringer. 1965. Pleistocene pollen analysis and biogeo- graphy of the Southwest, p. 433-451. In: H. E. Wright and D. G. Frey (eds.), The Quaternary of the United States, Princeton Univ. Press, Princeton, N.J. Miller, A. H. 1935. Some breeding birds of the Pine Forest Mountains, Nevada. Auk 52: 467-468. 1946. Vertebrate inhabitants of the pinyon association in the Death Valley region. Ecology 27: 54-60. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 227 Miller, A. H. and W. C. Russell. 1956. Distributional data on the birds of the White Mountains of California and Nevada. Condor 58: 75-77. Miller, R. R. 1948. The cyprinodont fishes of the Death Valley system of eastern California and southwestern Nevada. Misc. Publ. Univ. Mich- igan Mus. Zool. 68: 1-155. Power, D. M. 1972. Numbers of bird species on the California islands. Evolution 26: 451-463. Preston, F. W. 1962. The canonical distribution of commonness and rarity: Part I. Ecology 43: 185- 215. Schoener, A. 1974. Experimental zoogeography: colonization of marine mini-islands. Amer. Nat- uralist 108: 715-738. Seifert, R. P. 1975. Clumps of Heliconia in- florescences as ecological islands. Ecology 56: 1416-1422. Simberloff, D. S. 1974. Equilibrium theory of is- land biogeography and ecology. Ann. Rev. Ecol. Syst. 5: 161-182. Simberloff, D. S., and E. O. Wilson. 1969. Experimental zoogeography of islands. The colo- nization of empty islands. Ecology 50: 278-296. Simpson, R. R. 1974. Glacial migrations of plants: island biogeography evidence. Science 185: 689- 700. Smith, A. T. 1974a. The distribution and dispersal of pikas: consequences of insular population structure. Ecology 55: 1112-1119. 1974b. The distribution and dispersal of pikas: influence of behavior and climate. Ecolo- gy 55: 1368-1376. Smith, G. R. 1978. Riogeography of Intermountain fishes. Great Rasin Nat. Mem. 2: 17-42. Terborgh, J. 1973. Chance, habitat, and dispersal in the distribution of birds in the West Indies. Evolution 27: 338-349. Van Rossem, A. J. 1936. Rirds of the Charleston Mountains, Nevada. Pac. Coast Avif. 24: 1-65. Vuilleumier, F. 1970. Insular biogeography in conti- nental regions. I. The northern Andes of South America. Amer. Naturalist 104: 373-388. 1973. Insular biogeography in continental re- gions. II. Cave faunas from Tesin, southern Switzerland. Syst. Zool. 22: 64-76. Wells, P. V., and R. Rerger. 1967. Late Pleisto- cene history of coniferous woodland in the Mo- jave Desert. Science 155: 1640-1647. Wells, P. V., and C. D. Jorgensen. 1964. Pleistocene wood rat middens and climatic change in the Mojave Desert: a record of juni- per woodlands. Science 143: 1171-1173. West, N., R. J. Tausch, K. H. Rea, and P. T. Tueller. 1978. Phytogeographical variation within juniper-pinyon woodlands of the Great Rasin. Great Rasin Nat. Mem. 2: 119-136. BIOGEOGRAPHY AND MANAGEMENT OF NATIVE WESTERN SHRUBS: A CASE STUDY, SECTION TRIDENTATAE OF ARTEMISIA E. Durant McArthur and A. Perry Plummer1 Abstract.— Biogeographical considerations are important in the management of western shrublands. Seedings on rangelands have a higher probability of success when tried and tested principles are followed. It is usually best to seed mixtures that include adapted shrubs and herbs. Shrubs generally are well adapted to the environ- mental extremes of western ranges. Throughout the year, they provide nutrients for herbivores that are only sea- sonably available in herbs. Section Tridentatae of Artemisia is endemic to western North America and distinct from the analogous Eura- sian section (or subgenus) Seriphidium. The groups have separate, distinguishable centers of diversity; the two groups seem to be connected in the geologic past by way of the more primitive subgenus Artemisia. Preliminary karyotypic evidence suggests different but advanced karyotypes for both groups; chemotaxonomic and morpholo- gical data indicate differences in the groups. The Tridentatae likely evolved in North America during late Ter- tiary or early Quaternary times under the stimulus of cycles of aridity. The Tridentatae (sagebrushes) are morphologically variable. Different accessions are differentially adapted. Management practices for various taxa should take into consideration the individual taxon's characteristics. Effort should be made to seed adapted taxa and accessions. Sagebrush management requires maintenance, seeding, or thinning, depending upon the circumstances. Aside from serving as a review of the lit- erature, this paper presents original data and touches upon unpublished material and reports being prepared for publication. Shrubs are an important component of the vegetation of the American West, as well as in other parts of the world. In fact, much of the West is classified as shrubland because the dominant plant species are shrubs (Kiichler 1964, McGinnies 1972). Plummer (1974) recognized six major shrub- land types in the American West. Since the arrival of the white man, shrub ecosystems have been exploited and des- poiled by his livestock and land manage- ment practices (Cottan 1961, Heady 1975). The distribution and composition of many plant and animal communities have been substantially altered. The relatively new sci- ence of range management was born, in part, to systematize efforts to make shrub- lands more productive. The value of shrubs in their own right has long been under- estimated (McKell 1975b). In fact, much ef- fort, time, and money have been spent to eradicate and control shrubs in order to fa- cilitate the establishment of exotic grass monocultures. Shrubs have many current and potential uses. In the proceedings of an international symposium on useful wildland shrubs (McKell et al. 1972), a section on the pres- ent and possible uses of shrubs included chapters devoted to browse and cover for wildlife, low-maintenance landscaping, soil cover and stabilization, fire relations, medic- inal values, and industrial raw materials. This list was not exhaustive; for example, no reference is made to the use of shrubs as food for livestock. Our work on shrubs began with a need to improve critical winter game ranges (Plum- mer et al. 1968) and has continued with emphasis shifting to improving all types of disturbed sites (McArthur et al. 1974, Mon- sen 1975, Plummer, in press). Numerous 'Research geneticist and project leader, Intermountain Forest and Range Experiment Station, Forest Service, U.S. Department of Agriculture, Ogden, Utah 84401; located at the Intermountain Station's Shrub Sciences Laboratory, Frovo, Utah 84601. 229 230 GREAT BASIN NATURALIST MEMOIRS No. 2 species and ecotype adaptation plots have been set out both in and out of the natural ranges of particular shrub taxa. Plummer (in press) listed 85 native and 14 exotic shrubs that have proved to be adapted to disturbed sites in one or more vegetative communities of the Intermountain Region. Some shrubs can be established only vegetatively or by transplanting and do not reproduce by seed. Even these, however, have value since they cover and stabilize disturbed areas and serve as a nurse crop on critical sites for plants that establish more slowly. A principal thrust of range management has been in the modification of vegetation— often disclimax vegetation (Heady 1975). Mechanical, chemical, and burning methods have been used widely to control unwanted plants— mostly woody ones— on rangelands. These treatments are generally followed by seeding. Practical trials and experiments have succeeded and failed often enough for cause and effect relationships to be ana- lyzed and to serve as a set of principles. Recommendations relative to the seeding of many range sites can be made with a high probability of success. Plummer et al. (1968) and Heady (1975) outlined procedures and site criteria necessary for successful range seedings. Except for weather conditions, these are controllable by the range man- ager. We wish to emphasize, as did Plum- mer et al. (1968) and Heady (1975), the im- portance of planting mixtures of adapted taxa because: (1) Many seedings are on variable terrain that includes diverse microhabitats. (2) A mixed diet is usually more palatable to and nutritious for herbivores. (3) Periods of growth vary for different plant taxa and classes (shrubs, forbs and grasses); so succulent forage is provided for a longer time. (4) Some plants benefit others by provid- ing habitat and nutrients. (5) Diseases and insect pests do not at- tack all species equally. Shrubs should be included in most seedings since they are well adapted to drought, salinity, acidity, wide temperature fluctuations, and other environmental extremes of western ranges. In addition, they provide habitat for ani- mals that other classes of plants do not (McKell 1975a). This paper examines the biogeography and the management implications of a par- ticular group of shrubs— the sagebrushes, section Tridentatae of Artemisia. The sage- brushes are perhaps the most common shrub in western North America. In order to get a better idea of how to better manage and utilize this resource, we have tried to catch a glimmer of its evolutionary past. Such in- formation should benefit the management of sagebrush ranges by providing bases for bet- ter understanding species adaptation and distribution and plant improvement pro- grams through hybridization and selection. M ETHODS AND MATERIALS M Species Distributions.— General distribu- tion of species of the Tridentatae was ob- tained from Ward (1953), Reetle (1960), and by examining the Artemisia collections of the herbaria of Brigham Young University (BRY) and the Intermountain Station's Shrub Sciences Laboratory (SSLP). We have followed Beetle's (1960) and Beetle and Young's (1965) nomenclature for the Triden- tatae (Fig. 1). Distribution of the Seriphi- dium species that occurs in the Soviet Union was taken from Polyakov (1961). For spe- cies that occur outside the Soviet Union, de Candolle (1838) and Boissier (1875) were consulted (Fig. 2). We reviewed other re- gional and national floras peripheral to the Soviet Union but found little information to add to the aforementioned references. Karyotyping.— The karyotypes of section Tridentatae were obtained from a prelimi- nary analysis of data for a later report (McArthur, Pope, and Plummer, in prepara- tion). A sample cell was selected from each of 13 diploid and 14 tetraploid populations representing 9 diploid and 9 tetraploid taxa. Squashed root tips of seedlings were micro- scopically analyzed. The slides were pre- pared by fixing colchicine-pretreated root tips in 1:3 acetic alcohol and squashing them in acetocarmine. The idiogram of sec- tion Tridentatae (Fig. 3) was prepared from measurements made on 20 X 25 cm photo- micrographs prints at a magnification of 3120X. The idiogram for the two species of 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 231 126 120 Fig. 1. Distribution ranges of the species of section Tridentatae. 232 GREAT BASIN NATURALIST MEMOIRS No. 2 Seriphidium (Fig. 3) was prepared from data presented by Filatova (1974). We used the shorter of two lengths Filatova gives for chromosome 3 of A. juncea. Chromosome sizes (L = large, M = medium, S = small) were determined by pairing the chromo- somes according to relative length and cen- tromere position, dividing the genome into relative lengths, averaging the length of each pair, and proportionalizing the pair so the genome had a length of 100 arbitrary units. Thus, L > 12.4, M = 9.6 - 12.4, S < 9.6 for diploids and L > 6.2, M = 4.8 - 6.2, S < 4.8 for tetraploids. The centro- mere positions (M = metacentric, SM = submetacentric, and ST = subterminal) were determined by the ratio of the length of the short arm to the length of the long arm. Thus, M > 75 percent, SM = 50 - 75 percent, ST < 50 percent (Tables 1 and 2). Species Adaptation.— Shrub wildings have been collected in the spring (March- June) and fall (October-November) and trans- planted to uniform gardens and smaller spe- cies adaptation plots. Periodic ratings are made as to their height, crown, vigor, her- bage yield, reproduction, and survival. These data are on file principally at the Great Basin Experimental Area in Ephraim, Utah, but also at the Shrub Sciences Labo- ratory in Provo, Utah. Collection numbers, prefixed by a U, have been assigned to ac- cessions within each taxon (Tables 1, 2, 3). Voucher herbarium specimens are deposited at the Shrub Sciences Laboratory (SSLP). Section Tridentatae of Artemisia The genus Artemisia is a group of about 200 plant species which belong to the tribe Anthemideae of Compositae (de Candolle 1837, Clapham 1962). Good (1974) consid- ers Artemisia as belonging to his temperate genera, group 1, subgroup 2, which includes Fig. 2. Distribution of the section (subgenus) Seriphidium in Eurasia and North Africa. The numbers indicate maximum number of species in each area. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 233 "genera found throughout the northern ex- tratropical latitudes with some extension southwards in all directions, usually to cer- tain tropical mountains only." Besser (1829) began the first com- prehensive monograph on the genus early in the 19th century. He recognized four sec- tions— Abrotanum, Absinthium, Dracunculus, and Seriphidium. His monograph was not completed before he died, but de Candolle's (1837) "Prodomus Systematis Naturalis" and Hooker's (1840) "Flora Boreali-Americana" include significant portions of Besser's work. Besser's sections are still considered the principal divisions of the genus, but in some treatments they have been elevated to the rank of subgenus and additional subordinate sections created (Rydberg 1916, Polyakov 1961). The comprehensive treatment of Ar- temisia in the "Flora of the USSR." (Poly- akov 1961) unites Absinthium and Abrota- num into the single large subgenus Artemisia. Hall and Clements (1923) recog- nized the close affinity of these two groups, which were separated in Besser's scheme only by the hairiness of the floral recep- tacle. Members of Seriphidium are distin- guished by having only perfect disk flowers whereas members of the other subgenera have fertile or sterile ray flowers and vari- ous degrees of disk flower fertility. Rydberg (1916), while recognizing the subgenus Seri- phidium, created the section Tridentatae for most North American members of Seriphi- dium. He also created monotypic sections for A. rigicki and A. pygmaea as well as a new genus for the largely herbaceous A. palmeri. More recent monographers (Hall and Clements 1923, Ward 1953, and Beetle 1960) have been more conservative than Rydberg, but have been divided on the re- tention of Tridentatae as a section. Beetle (1960) included all woody, North American, homogamous Artemisia species in the sec- tion Tridentatae along with A. bigelovii, which often has a ray flower or two on oth- erwise discoid heads. Beetle's Tridentatae form a group of 11 unequal species endemic to western North America (Fig. 1). The spe- cies distribution ranges vary in size from SERIPHIDIUM TRIDENTATAE Fig. 3. Idiograms of two species of section Seriphidium, A. juncea, and A. leucodes (after Filatova 1974) and a generalized idiogram of section Tridentatae. Note secondary constriction in long arm of the first chromosome in the Tridentatae. The arrows indicate a chromosome substitution in some taxa (Table 1). 234 GREAT BASIN NATURALIST MEMOIRS No. 2 -r >n o o — ID IC 05 C ~ A £ 4*1 Z A w C ~ - — 3) ■3 ° 3 U '' $ - £ -9 J ^ «p g« 4 -S «A ° "9 • a 3 * - c - < < D 5685 - 1 .0496 X < scheme effectively predicts plant- and animal-based life zone boundaries, if one can ignore a few "indicator" species. These data reinforce the general findings of others mentioned previously that the Holdridge scheme shows promise, but are more per- suasive in that they use plants and animals in a topographically complex area where distributions are well known. An interesting offshoot of these analyses are shown in Figures 9 and 10. It is usually assumed that long-term weather records would be necessary to derive Holdridge's climatic variables. Our data suggest, for Utah, that mean annual biotemperature can be predicted from mean annual temperature (Fig. 9). This is not surprising, but it means that one can take mean annual temperature data from a weather station in a valley cor- rected for altitude, etc., to get a corrected mean annual temperature for a place dis- tant from that station. Mean biotemperature may be read directly from the curve or cal- culated from the formulae. Additionally, for Utah sites, a single measurement of snow accumulation on one day a year (1 April, Snow Water Equivalents) predicts annual precipitation (Fig. 10). For Utah this high Fig. 9. Relationship between mean annual biotem- perature and mean annual temperature for Utah weather stations. Fig. 10. Relationship between the 1 April snow wa- ter equivalent (SWE) and the annual precipitation for localities in the Manti-LaSal National Forest having snow courses studied by the United States Soil Con- servation Service. 256 GREAT BASIN NATURALIST MEMOIRS No. 2 correlation is due in part to the high per- centage of precipitation (60 percent or more) that comes as snowfall. A similar relationship may be of no value in other areas, but these results suggest that a person might be able to develop various other cor- rections specific to particular geographic areas. The prospect of using simple data ex- tends the potential use of the Holdridge scheme. Additionally, since one only needs values for two variables to assign an area to a life zone hexagon, one may actually read the approximate value of the unknown third parameter off of the life zone model (Thompson 1966). The Holdridge life zone model does pre- dict plant and animal aggregations for Utah with reasonable accuracy. The next step to reach the final goal of using this scheme to portion the earth into a reasonable number of subsections for management purposes is to apply the scheme to another area con- taining organisms not closely related to those of Utah and then to determine wheth- er or not a similar value of C obtains for the unrelated areas falling into the same life zone. That analysis must be left for another place and time. Acknowledgments This work was made possible by the US/IBP Desert Biome funded by the Na- tional Science Foundation (Grant GB32139). Donna Baranowski and Bette Peitersen aided in manuscript preparation, and Linda Finchum typed the manuscript. The United States Forest Service, Dixie National Forest, and Soil Conservation Service, Manti-LaSal Snow Course Project, kindly permitted use of their data. Literature Cited Reals, E. W. 1969. Vegetational change along al- titndinal gradients. Science 165: 981-985. Cronquist, A., A. H. Holmgren, N. H. Holmgren, and J. L. Reveal. 1972. Intermountain Flora: Vascular plants of the Intermountain West. Vol. 1 Hafner Publ. Co., New York. Diamond, J. M. 1975. The island dilemma: lessons of modern biogeographic studies for the design of natural reserves. Biol. Conserv. 7: 129-146. 1976. Island biogeography and conservation: strategy and limitations. Science 193: 1027-1029. Durrant, S. D. 1952. Mammals of Utah. Univ. Kan- sas Publ. Mus. Natl. Hist. 6: 1-549. Franklin, J. F. 1977. The biosphere reserve pro- gram in the United States. Science 195: 262- 267. Heyer, W. R. 1967. A herpetofaunal study of an ecological transect through the Cordillera de Tilaran, Costa Rica. Copeia 1967: 259-271. Holdridge, L. R. 1947. Determination of world plant formations from simple climatic data. Sci- ence 105: 367-368. 1967. Life zone ecology. Rev. ed. Tropical Science Center, San Jose, Costa Rica. Holdridge, L. R., W. C. Grenke, W. H. Hatheway, T. Liang, and J. A. Tosi. 1971. Forest envi- ronments in tropical life zones— a pilot study. Pergamon, New York. Jeppson, R. W., G. L. Ashcroft, A. L. Hurer, G. V. Skogerboe, and J. M. Bagley. 1968. Hydrologic atlas of Utah. Utah Water Research Laboratory, Utah State University, Logan. Kuchler, A. W. 1964. Manual to accompany the map. Potential natural vegetation of the counterminous United States. Amer. Geogr. Soc. Spec. Publ. 36: 1-116. Lindsey, A. A., and J. O. Sawyer. 1970. Vegetation-climate relationship in the eastern United States. Proc. Indiana Acad. Sci. 80: 210- 214. Little, E. L. 1971. Atlas of United States trees. Vol. 1, conifers and important hardwoods. USDA Forest Service Misc. Publ. 1146. MacArthur, R. H., and E. O. Wilson. 1967. The theory of island biogeography. Princeton Monogr. Popul. Biol. 1: 1-203. MacMahon, J. A. In press-a. Thoughts on the opti- mum size of natural reserves based on ecologi- cal principles. In: J. F. Franklin (ed.), Biospher- ic reserves. USDA. MacMahon, J. A. In press-b. North American deserts. In: R. Perry (ed.), Deserts of the world. Cambridge University Press, Cambridge. Myers, N. 1976. An expanded approach to the problem of disappearing species. Science 193: 198-202. Preston, F. W. 1962. The canonical distribution of commonness and rarity. Ecology 43: 185-215. Price, R., and R. B. Evans. 1937. Climate of the west front of the Wasatch Plateau in central Utah. Monthly Weather Rev. 65: 291-301. Sawyer, J. O., and A. A. Lindsey. 1963. The Hold- ridge bioclimatic formations of the eastern and central United States. Proc. Indiana Acad. Sci. 73: 105-112. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 257 1971. Vegetation of the life zones in Costa Rica. Indiana Acad. Sci. Monogr. 2: 1-214. Simberloff, D. S. 1976. Species turnover and equi- librium island biogeography. Science 194: 572- 578. Simberloff, D. S., and L. G. Abele. 1976. Island biogeography theory and conservation practice. Science 194: 285-286. Slud, P. 1960. The birds of finca "La Selva", Costa Rica: a tropical wet forest locality. Bull. Amer. Mus. Natl. Hist. 121: 51-148. 1976. Geographic and climatic relationships of avifaunas with special references to com- parative distribution in the Neotropics. Smithso- nian Contr. Zool. 212: 1-149. Sullivan, A. L., and M. L. Shaffer. 1975. Biogeography of the megazoo. Science 189: 12- 17. Stebbins, R. L. 1966. A field guide to western rep- tiles and amphibians. Houghton-Mifflin, Boston. Terborgh, J. 1974. Preservation of natural diversity: the problem of extinction prone species. Bio- Science 24: 715-722. Thompson, P. T. 1966. A test of the Holdridge model in midlatitude mountains. Prof. Geogr. 18: 286-292. Tosi, J. A., Jr. 1964. Climatic control of terrestrial ecosystems: a report on the Holdridge model. Econ. Geogr. 40: 173-181. Wilson, E. O., and E. O. Willis. 1975. Applied biogeography, pp. 522-534. In: M. L. Cody, and J. M. Diamond (eds.), Ecology and evolution of communities. Harvard University Press, Cam- bridge. 1978 INTERMOl NTAIN BIOGEOGRAPHY: A SYMPOSIUM 259 SUBJECT AND SPECIES INDICES FOR INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM Abajo Mountains, p. 61, 77. Albion Mountains, p. 83. Allopatric ranges, p. 71, 150, 153, 201. Alpine biota, p. 11, 67, 72, 105, 109, 112. Altitude, p. 29, 245. Amphibians, p. 43, 245. Angiosperms, p. 5, 6, 7. Annual plants, p. 85, 87, 115, 170. Aquarius Plateau, p. 65. Arctic, p. 11, 105, 111, 112. Arcto-Tertiary flora, p. 8, 9, 10, 238. Area, p. 90, 91, 122, 209, 220, 245. Arizona, p. 4, 43, 68, 105. Ash Meadows, p. 140. Asia, p. 7, 8, 237. Aspen forest, p. 137. Australia, p. 8. Barriers, p. 17, 20, 28, 29, 30, 31, 32, 38, 47, 61, 63, 66, 71, 83, 84, 88, 107, 121, 122, 152, 167, 197, 217. Baseline data, p. 72. Beartooth Mountains, p. 105, 110, 111. Beaver Dam Mountains, p. 50, 58, 73. Bees, p. 13. Beringia, p. 10, 13, 237. Biogeoraphie regions, p. 106, 115, 119, 120, 134, 137, 202, 245. Biogeography, p., 158, 209, 245. Biotemperature, p. 255. Bird dispersed plants, p. 96, 97. Birds, p. 5, 12, 55, 89, 98, 137, 209, 220. Bison, p. 12. Black Pine Mountains, p. 122. Blue Lake, p. 101. Body size, p. 68, 69, 99, 215, 221. Boreal biota, p. 31, 55, 59, 63, 71, 77, 137, 157, 209. Bryce Canyon National Park, p. 83, 88, 93. Burbank Hills, p. 122. Burro, p. 12. California, p. 13, 43, 55, 56, 57, 82. Camel, p. 12. Canada, p., 11, 161. Carnivores, p. 222. Carson Range, p. 212. Cassia Mountains, p. 83. Center of distribution, p. 7, 8, 237. Charleston Peak, p. 105. Chemotaxonomy, p. 238. Chihuahua, p. 46. Chile, p. 8. Circumboreal flora, p. 11. Clark Mountain, p. 98. Climate, p. 3, 85, 88, 99, 108, 119, 130, 134, 144, 197, 245. Clinal variation, p. 55, 64, 65, 66, 67, 68. Coahuila, p. 46. Colonization, p. 17, 31. Color, p. 68. Colorado, p. 4, 6, 44, 56, 57, 67, 68. Colorado Plateau, p. 4, 43, 47, 57, 192, 198, 206, 255. Colorado River, p. 17, 20, 27, 31, 33, 34, 38, 67, 139, 140. Community structure, p. 141. Competition, p. 37, 129, 152. Confusion Range, p. 122. Coniferous forest, p. 11, 137, 147, 221. Continental climate, p. 3, 4, 7, 99. Continental drift, p. 4, 18. Craters of the Moon, p. 5. Cretaceous, p. 4, 5, 6, 7, 9, 12, 105. Death Valley, p. 20, 26, 32. Deciduous trees, p. 8. Deep Creek Mountains, p. 61, 65, 68, 72, 73, 77, 83, 92, 93, 111, 212. Desatoya Range, p. 212. Diamond Mountains, p. 212. Dicotyledons, p. 6, 7. Dinosaurs, p. 12. Diptera, p. 13. Disjunct ranges, p. 44, 81, 147, 153, 237. Dispersal, p. 5, 17, 27, 71, 81, 93, 94, 96, 98, 111, 197, 199, 212. Distance, p. 89, 90, 91, 93, 114, 200, 209. Drainage basins, p. 18, 19, 20, 21. East Humboldt Mountains, p. 108, 122, 133. East Tintic Mountains, p. 83, 93. Ecological role, p. 141. Edaphic factors, p. 10, 89, 99, 109, 110, 260 GREAT BASIN NATURALIST MEMOIRS No. 2 115, 116, 161, 164, 191, 197, 205, 207. Elephant, p. 12. Elevation, p. 11, 83, 85, 89, 90, 91, 122, 126, 220. Elk Ridge, p. 61. Elko, p. 140. Ely, p. 140. Endangered species, p. 24, 25, 37, 72. Endemics, p. 13, 17, 22, 35, 36, 43, 55, 57, 81, 87, 91, 97, 105, 110, 111, 113, 171, 192, 204, 229. Eocene, p. 4, 5, 6, 7, 12, 18. Eurasia, p. 7, 8, 9, 12. Europe, p. 13. Evolution, p. 3, 8, 13, 36, 115, 161, 170, 229, 238. Excelsior Mountains, p. 122. Extinct species, p. 22, 24, 25, 98, 217, 221. Extinction, p. 3, 12, 13, 17, 28, 31 32, 37, 39, 73, 98, 99, 209. Fauna, p. 3. Fidelity index, p. 199. Fish, p. 17, 210, 223. Fleshy fruits, p. 84, 94, 96. Flora, p. 3, 81, 86. Forbs, p. 96. Fossils, p. 17, 25, 26, 28, 44, 55, 56, 166. Frenchman Flat, p. 43. Frisco Mountains, p. 61, 77. Geologic history, p. 3, 4, 71. Glaciers, p. 10, 20, 44, 69, 71, 106, 108, 161, 192. Gondwanaland, p. 6, 9, 13. Goose Creek Mountains, p. 122. Goose Lake, p. 20, 33, 34. Graminoides, p. 87, 96. Grand Canyon, p. 17, 33. Grant Range, p. 212. Grasses, p. 3, 87, 96. Grassland, p. 10, 73, 170. Great Rasin, p. 3, 17, 19, 43, 45, 47, 49, 55, 57, 81, 119, 198, 209, 255. Great Plains, p. 44, 55. Great Salt Lake, p. 55. Green River, p. 27. Green River flora, p. 6. Gymnosperms, p. 5. Habitat diversity, p. 59, 63, 89, 119, 122, 125, 147, 197, 209, 215, 220. Harney Rasin, p. 34. Henry Mountains, p. 61, 77. Herbaceous plants, p. 3, 7, 8, 98, 170. Herbivores, p. 222. Highland Mountains, p. 122. Holdridge's Life Zones, p. 245, 246, 247. Horse, p. 12. House Range, p. 61, 77. Humboldt River, p. 139, 140. Hybridization, p. 8, 10, 11, 55, 58, 69, 153, 156, 161, 241. Hymenoptera, p. 13. Hypsithermal (Altithermal) period, p. 11, 20, 109, 111, 129, 167. Idaho, p. 4, 43, 44, 56, 82, 140. Immigration, p. 209. Independence Mountains, p. 108. Insects, p. 13. Intercalary meristem, p. 12. Introductions, p. 17, 37, 39, 72, 98. "Islands," p. 5, 17, 55, 59, 72, 73, 81, 82, 88, 90, 105, 106, 161, 209. Jarbidge Mountains, p. 72, 83, 97 108, 115. Jemez Mountains, p. 44. Jurassic, p. 5, 12, 13. Kaibab Plateau, p. 83. Karyotype, p. 161, 229, 230, 231, 238. Kawich, p. 140. Klamath River, p. 17, 20, 26, 33, 34. Lake Ronneville, p. 20, 26, 32, 34, 71, 161. Lake Idaho, p. 19, 24. Lake Lahontan, p. 20, 26, 32, 34, 55, 71, 131, 161. La Sal Mountains, p. 61, 77. Lassen Volcanic National Park, p. 83. Latitudinal effects, p. 126, 137, 142. Laurasia, p. 6, 9, 13. Lava, p. 4. Lifeform of plants, p. 84, 87. Longevity, p. 99. Longitudinal effects, p. 126. Madro-Tertiary flora, p. 8, 9, 10. Mainlands (continents), p. 88, 90, 121, 215. Mammals, p. 5, 11, 31, 55, 62, 89, 98, 209, 220, 245. Management, p. 1, 2, 17, 39, 53, 72, 100, 101, 134, 229, 230, 239, 245. Markagunt Plateau, p. 65. Meadow Valley Wash, p. 38, 58, 139, 140. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 261 Mediterranean region, p. 8. Merriam effect, p. 129. Mexico, p. 7, 8, 9, 43, 46, 161. Mineral Mountains, p. 61, 77, 122. Miocene, p. 3, 4, 7, 10, 12. Mojave biota, p. 202. Mojave Desert, p. 7, 13, 38, 55, 58, 105, 125, 253. Molluscs, p. 55. Monitor Range, p. 122, 133. Monocotyledons, p. 7. Montana, p. 7, 68. Montane biota, p. 11, 63, 203, 212. Mount Timpanogos, p. 83, 93. Nature reserves, p. 245. Navajo Mountains, p. 61, 77. Needle Range, p. 61, 77, 122. Nevada, p. 4, 18, 43, 45, 55, 56, 65, 67. New Mexico, p. 4, 43, 44, 68. New World, p. 9. Niche breadth, p. 28, 61, 96, 100, 200, 207, 222. Normalizing selection, p. 201. North America, p. 7, 8, 9, 12. Oak woodland, p. 10, 11, 129, 137. Olancha Peak, p. 105, 111. Old World, p. 7, 8. Oligocene, p. 4, 6, 7, 18, 105. Oquirrh Mountains, p. 61, 65, 66, 68, 77, 212. Oregon, p. 33, 34. Overgrazing, p. 53. Overkill, p. 3. Owens River, p. 20, 26, 32. Pack rat middens, p. 43. Pahranagat, p. 140. Pahrump, p. 140. Paleocene, p. 6, 7, 105. Palm, p. 5, 6. Panamint Range, p. 212. Parasites, p. 99. Paunsaugunt Plateau, p. 65, 212. Pavant Mountains, p. 65. Permian, p. 13, 105. Physiological adaptations, p. 116. Phytogeography, p. 105. Pilot Peak, p. 122, 212. Pine Forest Mountains, p. 97, 101. Pine Valley Mountains, p. 61, 77, 83, 93, 122, 198. Pinyon-juniper woodland, p. 10, 43, 55, 57, 73, 100, 119, 137, 142, 213. Pit River, p. 17. Plant succession, p. 133. Pleistocene, p. 3, 12, 44, 48, 55, 63, 71, 81, 221, 223. Pleistocene pluvials, p. 17, 19, 43, 105. Pliocene, p. 4, 7, 17, 38, 105. Pluvials, p. 31, 71, 166. Pollen record, p. 5, 240. Pollinating insects, p. 3, 13. Polyploidy, p. 10, 109, 240. Population size p. 221. Precipitation, p. 255. Precipitation/evaporation ratio, p. 5, 10, 11, 28, 31, 71. Predation, p. 37. Primitive man, p. 3, 55. Pyramid Lake, p. 55. Quaternary, p. 56, 170. Raft River Mountains, p. 61, 67, 68, 77, 83. Range limits, p. 197. Red Butte Canyon, p. 83, 93. Refugia, p. 43, 48, 115. Reptiles, p. 43, 245. Riparian woodland, p. 137, 147, 205. Roberts Creek Mountains, p. 212. Rocky Mountains, p. 4, 11, 13, 17, 55, 99, 105, 108, 215. Ruby Mountains, p. 20, 33, 72, 83, 87, 92, 93, 97, 105, 108, 110, 111, 115, 133, 212. Sacramento Mountains, p. 44. Sagebrush steppe, p. 55, 56, 73, 129. Sagehen Creek, p. 83. San Francisco Peak, p. 105, 108, 111, 113, 115. Santa Rosa Mountains, p. 83, 97, 108. Saprophytes, p. 99. Schell Creek Mountains, p. 122, 212. Sequoia-Kings Canyon National Park, p. 83. Sevier River, p. 53. Sheep Mountains, p. 212. Shoshone Mountains, p. 122. Shrubs, p. 8, 87, 96, 170, 229. Sierra Nevada Mountains, p. 4, 17, 31, 55, 57, 67, 81, 96, 99, 105, 108, 111, 212, 215. Silene acaulis, p. 110, 112. 262 GREAT BASIN NATURALIST MEMOIRS No. 2 Similarity indices, p. 17, 21, 26, 27, 95, 105, 113. Snake Range, p. 72, 109, 212. Snake River, p. 17, 20, 26, 31, 32, 33, 139. Snake River Plain, p. 4, 17, 18, 38, 44. Sonora, p. 46. Source areas, p. 81, 90, 91, 92, 93, 115, 141, 164. South Africa, p. 8. South America, p. 8, 9, 12. Southwest, p. 57. Speciation, p. 138, 152. Species-area curves, p. 17, 31 61, 81 88, 97, 122, 215, 245. Species diversity, p. 31, 37, 56, 59, 81, 82, 87, 96, 101, 105, 122, 124, 125, 138, 139, 144, 220. Species equilibrium, p. 37, 81, 98, 99, 134, 209, 211. Spring Range, p. 83, 87, 91, 108, 111, 212. Spruce Mountain, p. 133, 212. Stansbury Mountains, p. 61, 77, 212. Sticktights, p. 84, 94, 96. Stream Capture, p. 27. Subalpine biota, p. 67. Sulphur Springs Range, p. 133. Sympatric species, p. 36, 71, 150, 153. Synecology, p. 119. Tavaputs Plateau, p. 65. Temperature inversions, p. 119, 131. Tertiary, p. 4, 6, 7, 8, 9, 12. Teton Mountains, p. 81. Therapsid reptiles, p. 12. Threatened species, p. 17, 23, 25. Timberline, p. 11, 106. Toana Mountains, p. 122. Toiyabe Mountains, p. 83, 87, 93, 108, 110, 111, 122, 133, 140, 212. Toquima Range, p. 110, 212. Transition zone, p. 197. Trees, p. 7, 87, 96, 99. Triassic, p. 12, 105. Trophic dynamics, p. 81. Truckee River, p. 140. Turnover rates, p. 223. Turtle Range, p. 98. Tushar Mountains, p. 65, 67, 122. Uinta Mountains, p. 44, 61, 64, 66, 68, 83, 212. Ungulates, p. 12. Utah, p. 4, 6, 43, 50, 51, 55, 56, 67, 82, 245. Utah Lake, p. 31. Vagility, p. 209, 223. Vegetation, p. 3, 72, 81, 119, 245. Vicariant species, p. 17, 27, 36. Virgin River, p. 33, 38, 47, 58, 68, 198. Volcanic activity, p. 17, 18. Vulnerable species, p. 17, 23, 24, 25. Wah Wah Mountains, p. 61, 77. Warner Mountains, p. 71, 83. Wasatch Mountains, p. 4, 17, 19, 31, 38, 44, 55, 61, 64, 66, 68, 77, 81, 83, 93, 96. Washington Co., Utah, p. 197. Wheeler Peak, p. 72, 83, 87, 93. White Mountains, p. 83, 108, 111, 113, 122, 212. White Pine Mountains, p. 212. White River, p. 33, 38. Wind dispersal, p. 84, 94, 96. Wyoming, p. 4, 43, 82. Yellowstone region, p. 18, 83. Yosemite National Park, p. 212. Z-values, p. 17, 31, 89, 209, 216. Zion Canyon, p. 58. Abies bahamea, p. 11. Abies concolor, p. 97, 98, 100, 148. Abies lasiocarpa, p. 11, 86, 100. Accipiter cooperi, p. 77. Accipiter gentilis, p. 11. Accipiter striatus, p. 77. Acer grandidentatum, p. 86, 100. Acipenser transmontanus, p. 22. Acipenseridae, p. 22. Aconitum columbianum, p. 86. Acrocheilus alutaceus, 23. Aegolius acadicus, p. 77. Aeroncmtes saxatlis, p. 77. Agastacke pallidiflora, p. 87. Agelaius phoeniceus, p. 67. Agropyron pringlei, p. 86. Agropyron spicatum, p. 92. AUenrolfca, p. 161. Allium, p. 11, 86. Amelanchier alnifolia, p. 92. Amelanchicr utahensis, p. 92. Androsace septentrionalis, p. 112. 1978 INTKHMOl NTAIN BIOOKOCHAPHY: A SYMPOSIUM 263 Aneides Jiardyi, p. 44. Angelica seabrida, p. 87. Antennaria marginata, p. 87. Antcnnaria rosea, p. 111. Antennaria soliceps, p. 87. Anthemidae, p. 7. Anthus spinoletta, p. 79. Aphelecora eaerulescens, p. 66. Aphelocoma coerulescens, p. 158. Aplodontia rufa, p. 225. Aquila chrysaetos, p. 77. Aquilegia triternata, p. 87. Archilochus alexandri, p. 78, 140. Archoplites, p. 26. Arctium, p. 85. Arctostaphylos patula, p. 94. Arenaria, p. 11. Arenaria confusa, p. 87. Arizona elegans, p. 48. Artemisia, p. 7, 229. Artemisia arbuscula, p. 92. Artemisia douglasiana, p. 86. Artemisia (section Seriphidium), p. 229. Artemisia (section Tridentatae), p. 229. Artemisia argilosa, p. 235. Artemisia bigelovii, p. 233, 238, 241. Artemisia cana, p. 235, 239. Artemisia junceus, p. 237. Artemisia leueodes, p. 237. Artemisia hngiloba, p. 235. Artemisia maritima, p. 237. Artemisia mendozana, p. 237. Artemisia nova, p. 239. Artemisia palmeri, p. 237, 239. Artemisia pygmaea, p. 233. Artemisia rigida, p. 233. Artemisia rothrockii, p. 235. Artemisia santonicum, p. 241. Artemisia tridentata, p. 7, 92, 100, 235, 239, 249. Asio otus, p. 77. Asteraceae, p. 7. A.stereae, p. 7, 8. Astragalus, p. 10. Atriplex, p. 8, 161. Atriplex acanthocarpa, p. 164. Atriplex canescens, p. 161, 163, 164, 241. Atriplex con ferti folia, p. 161, 163, 249. Atriplex corrugata, p. 161, 164. Atriplex cuneata, p. 161, 163. A t rip lex fa lea ta, p . 163. Atriplex gardneri, p. 161, 164 Atriplex garretH, p. 162, 164. Atriplex hymenelytra, p. 164. Atriplex navapensis, p. 162. Atriplex obovata, p. 161, 164. Atriplex polycarpa, p. 164. Atriplex torreyi, p. 164. Atriplex tridentata, p. 162. Balsamorhiza, p. 7, 86. Balsamorhiza macrophylla, p. 86. Balsamorhiza sagittata, p. 86. Bidens, p. 85. Bonasa umbeuus, p. 77. Boraginaceae, p. 8. Brassicaceae, p. 8. Bromus anomalies, p. 93. Bromus breviaristatus, p. 86. Bubo virginianus, p. 66, 77. Ba/o fcoreas, p. 49, 50, 51. Bufo microseaphus, p. 45. Bafo punctatus, p. 45. Bwfo ivoodhousei, p. 45. Bufeo jamaicensis, 66, 77. Calamagrostis scopulorum, p. 86. Callisaurus draconoides, p. 46. Caltha leptosepala, p. 93, 110. Calyptridium umbellatum, p. 111. Cardamine, p. 8. Carduelis pinus, p. 79. Care.r amplifolia, p. 86. Carex aurea, p. 86. Carex elynoides, p. 110. Carex helleri, p. 111. Carex lanuginosa, p. 86. Carex pulvinata, p. 110. Carex scopulorum, p. 110. Carex tahoensis, p. 86. Carpodaeus cassinii, p. 79. Castilleja clokeyi, p. 87. Castilleja linoides, p. 87. Castilleja viscidula, p. 87. Catharus fuscescens, p. 79, 140. Catharus ustulatus, p. 79, 140. Catostomidae, p. 22. Catostomus ardens, p. 24, 36. Catostomus catostomus, p. 24. Catostomus clarki, p. 24, 36. 264 GREAT BASIN NATURALIST MEMOIRS No. 2 Catostomus columbianus, p. 24. Catostomus discobolus, p. 24, 36. Catostomus fumeiventris, p. 24, 36. Catostomus insignis, p. 24. Catostomus latipinnis, p. 24. Catostomus luxatus, p. 25. Catostomus macrocheilus, p. 24, 36. Catostomus occidentalis, p. 24. Catostomus platyhrynchus, p. 17, 25, 27. Catostomus sp., p. 24. Catostomus tahoensis, p. 24, 36. Catostomus warnerensis, p. 24, 36. Ceanothus martini, p. 86, 93. Ceanothus velutinus, p. 93. Centurus uropygialis, p 140. Ceratodes, p. 10, 161. Certhia familiaris, p. 65, 78, 144, 148. Chaenactis, p. 7. Chamacbatiaria millefolium, p. 100. Charina bottae, p. 49, 50, 51. Chasmistes brevirostris, p. 25, 33. Chasmistes cirius, p. 25. Chasmistes liorus, p. 25. Chasmistes sp. p. 25. Cheilanthes gracillima, p. 86. Chenopodiaceae, p. 8, 161. Chordeiles minor, p. 67. Chlorocrambe hastata, p. 86. Chorizanthe, p. 170. Chrysothamnus, p. 10. Circaea, p. 85. Cirsium clokeyi, p. 87. Clematis columbiana, p. 86. Clethrionomys gapperi, p. 225. Cnemidophorus tigris, p. 48. Coccyzus americanus, p. 140. Coccyzus erythropthalmus, p. 140. Colaptes auratus, p. 67, 140. Colaptes cafer, p. 69. Colaptes chrysoides, p. 67. Coleogyne, p. 43. Coleonyx varigatus, p. 46. Coleoptera, p. 13. Coluber constritor, p. 49. Columba fasciata, p. 77. Contopus borealis, p. 78. Contopus sordidulus, p. 78. Corallorhiza, p. 99. Cottidae, p. 22. Cottus bairdn, p. 17, 25, 27, 29, 36. Cottus beldingi, p. 25. Cottus eonfusus, p. 25. Cottus echinatus, p. 25, 36. Cottus extensus, p. 25, 36. Cottus greenei, p. 26. Cottus klamathensis, p. 26. Cottus leiopomus, p. 26. Cottus pitensis, p. 26. Cottus princeps, p. 26. Cottus rhotheus, p. 26. Cottus tenuis, p. 26. Crenichthys baileyi, p. 25, 36. Crenichthys nevadae, p. 25, 36. Crotalus cerastes, p. 48, 50. Crotalus michelli, p. 48, 50. Crotalus scutulatus, p. 48, 50. Crotaphytus wislizeni, p. 46. Cryptantha, p. 8. Cryptantha mohavensis, p. 86. Cyanocitta stelleri, p. 66, 78, 144, 157. Cycloloma atripticifolium, p. 162. Cymopterus nivalis, p. 87. Cyprinidae, p. 22. Cyprinodon breviradis, p. 25, 36. Cyprinodon diabolis, p. 25. Cyprinodon nevadensis, p. 25. Cyprinodon radiosus, p. 25. Cyprinodontidae, p. 22. Cystopteris fragilis, p. 112, 115. Dedeckera, p. 170. Delphinium occidcntale, p. 93. Dendragnpus obscurus, p. 65, 77, 157. Dendrocopos pubescens, p. 140, 149, 157. Dendrocopos scalaris, p. 140. Dendrocopos villosus, p. 140. Dendroica coronata, p. 79. Dendroica gracial, p. 79, 148. Deschampsia caespitosa, p. 110. Dipsosaurus dorsalis, p. 46. Dodecatheon jeffreyi, p. 111. Dw/w. p. 110. Draba arida, p. 87. Draba lemmonii, p. 111. Dryocopus pileatus, p. 78. Dumetella carolinensis, p. 140. Eleocharis montana, p. 87. Ernpetrichthys lotos, p. 25, 36. 1978 INTERMOUNTAIN BIOGEOCRAPHY: A SYMPOSIUM 265 Empetrichthys merriami, p. 25, 36. Empidonax difficilis, p. 78, 147, 150, 157. Empidonax hammondii, p. 78, 147, 150. Empidonax minimus, p. 152. Empidonax oberholseri, p. 78. Enceliopsis, p. 7. Epilobium angustifolium, p. 86. Eremichthys acros, p. 23. Eremophila alpestris, p. 67, 78. Erigeron, p. 8, 11. Erigeron peregrinus, p. 110. Erigeron ursinus, p. 86. Erigeron ivatsoni, p. 87. Eriogoneae, p. 186. Eriogonoideae, p. 170. Eriogonum, p. 10, 11, 110, 170. Eriogonum gracilipes, p. 110. Eriogonum hohngrenii, p. 87. Eriogonum kingii, p. 87. Eriophyllum, p. 7. Eumeces skiltonianus, p. 50. Eutamias alpinus, p. 225. Eutamias amoenus, p. 225. Eutamias dorsalis, p. 100, 221, 225. Eutamias panamintinus, p. 225. Eutamias quadrimaculatus, p. 225. Eutamias quadrivittatus, p. 225. Eutamias speciosus, p. 225. Eutamias townsendii, p. 225. Eutamias umbrinus, p. 100, 221, 225. Ffl/co sparverius, p. 77, 140. Festuca arizonica, p. 87. Fritillaria atropurpurea, p. 86. Fundulus spp., p. 25. Gentiana, p. 11. Geothlypis tohniei, p. 79. Geothlypis trichas, p. 67. Geranium fremontii, p. 93. Geum macrophijllum, p. 86. Geum rossii, p. 86, 110. Gi/fl alvordensis, p. 23, 36. Gifa atraria, p. 23, 36. Gifo fcico/or, p. 17, 23, 27, 28,33, 36. Gi/a coerulea, p. 23. Gi/a copei, p. 23. Gt7a cj/p/ia, p. 23, 33. Gi/a elegans, p. 23, 33. Gi'/ia, p. 8. Gilmania, p. 170. Glaucidim gnoma, p. 77. Glaucomys sabrinus, p. 225. Glyceria elata, p. 86. Goodmania, p. 170. Gopherus agassizi, p. 46. Grazia, p. 161. Guiraca caerulae, p. 140. Hackelia floribunda, p. 86. Haplopappus, p. 7, 10. Harfordia, p. 170. Heliantheae, p. 7, 10. Hesperiphona vespertina, p. 148. Hesperoleucus symmetricus, p. 23. Hierochloe odorata, p. 86. Holbrookia maculata, p. 47. Hollisteria, p. 170. Holodiscus dumosus, p. 93. Hu/sea, p. 11, 86. Hidsea algida, p. 111. Hydrophyllaceae, p. 8. Ht//a arenicolor, p. 45. Hymenoxys, p. 11. Icterus bullockii, p. 69. Icterus cucullatus, p. 140. Icterus galbula, p. 69, 140. Icterus galbula, p. 140. Iotichthys phlegethontis, p. 23. Iridoprocne bicolor, p. 140. Ivesia pygmaea, p. 111. Junco caniceps, p. 70, 80, 157. Junco hycmalis, p. 70, 80, 157. Juniperus, p. 7, 85. Juniperus osteosperma, p. 43, 93, 98, 119. Kalmia polifolia, p. 94. Koeleria cristata, p. 110. Koenigia islandica, p. 112. Lampetra lethophaga, p. 22. Lampetra minima, p. 22. Lampetra tridentata, p. 22. Lampropeltis getulus, p. 48, 49. La Rivers, I., p. 18, 22, 23. Larrea, p. 4, 43. Lastarriea, p. 170. Lathyrus pauciflorus, p. 93. Ledum glandidosum, p. 94. Lepidium, p. 8. Lepidomeda albivallis, p. 24, 36. Lepidomeda altivelis, p. 24, 36. 266 GREAT BASIN NATURALIST MEMOIRS No. 2 Lepidomeda mollispinis, p. 24, 36. Leipidomeda nittata, p. 24. Leptotyphlops humilis, p. 48. Lepus americanus, p. 225. Lepus townsendii, p. 221, 225. Lesquerella, p. 8. Leucosticte atrata, p. 79. Leucosticte tephrocotis, p. 157. Lewisia pygmaea, p. 111. Lewisia rediviva, p. 93. Libocedrus decurrens, p. 86. Lithospermum, p. 8. Lomatium, p. 11. Lonicera involucrata, p. 86. Lo.ri'fl curvirostra, p. 79. Lupinus breweri, p. 111. Marmota flaviventris, p. 221, 225. Maries americana, p. 225. Masticophis flagellum, p. 48. Melospiza melodia, p. 140. Mertensia arizonica, p. 86. Mertensia toyabensis, p. 87. Microtus longicaudus, p. 221, 225. Microtus richardsoni, p. 225. Mimulus torreyi, p. 86. Moapa coriacea, p. 24. Moldavica parviflora, p. 86. Molothrus atcr, p. 66. Mucronca, p. 170. Muhlenbergia wrightii, p. 87. Mustela erminea, p. 221, 225. Myadestes townsendi, p. 79. Mylocheilus caurinus, p. 24. Mylopharodon, p. 26. Myosotis, p. 11. Myoxocephalus, p. 26. Nemacaulis, p. 170. Neotoma cinerea, p. 221, 225. Scotoma lepida, p. 43. Nicotiana attenuata, p. 94. Nucifraga columbiana, p. 78, 97, 98. Ochotona princeps, p. 221, 224, 225. Oenothera caespitosa, p. 93. Oncorhyncus tshawytscha, p. 22. Opheodrys oemalis, p. 49. Ophisaurus attenuatus, p. 44. Opuntia charlestonensis, p. 87. Oporonis tolmiei, p. 140. ( hobanche, p. 99. Orthocarpus tolmiei, p. 86. Orthodon, p. 26. Oryzopsis kingii, p. 86. Osmorhiza chilensis, p. 86. Otws as/o, p. 67, 140. Ofus flammeohis, p. 77. Oxynrt rfigi/na, p. 110, 111, 112, 115. Oxytheca, p. 170. Pachistima myrsinites, p. 93. Farws atricapillus, p. 66, 78, 140. P«n/.s gambeli, p. 65, 78. Passerella iliaca, p. 71, 140, 153. Passerina amoena, p. 69. Passerenia cyanea, p. 69. Pedicularis, p. 11. Penstemon, p. 10, 11. Penstemon keckii, p. 87. Peraphyllum ramosissirrium, p. 94. Perisoreus canadensis, p. 78. Petrochelidon pyrrhonota, p. 66. Phacelia, p. 8. Phainopepla nitens, p. 140. Phenacomys intermedins, p. 225. Plicucticus melanocephalus, p. 79. Phippsia algida, p. 112. P/i/o.v caespitosa, p. 111. P/i/o.t covillei, p. 110. P/i/o.v longifolia, p. 93. Phrynosoma douglassi, p. 50. Phijsaria, p. 8. Picea engelmannii, p. 11, 97, 100. Picea glauca, p. 11. P/refl pungens, p. 11, 86, 100. Picoides tridactylus, p. 78. Picoides villosus, p. 66, 78. Pinicola enucleator, p. 79. Pinus albicaulis, p. 94, 97. PinMS aristata, p. 94, 97, 98. Finns banksiana, p. 11. Pm?/.s contorta, p. 11. P/m/.s edufcs, p. 86, 97. Pinus flcxilis, p. 97, 98. Pinus jefferyi, p, 86, 98. Pnn/.s longaeoa, p. 109, 148. /'/M//.s monophylla, p. 98, 129. /'/m/.s ponderosa, p, 86, 94, 98, 100, 129, 148. /'//)//(> flfoertt, p. 140. Pipilo chlorura, p. 79. 1978 INTERMOUNTAIN BIOGEOGRAPHY: A SYMPOSIUM 267 Pipilo erythrophthalmus, p. 79. Piranga ludoviciana, p. 79. Piranga rubra, p. 140. Plagiobothrys, p. 8. Plagopterus argentissimus, p. 24. Plethodon neomexicanus, p. 44. Plethodontidae, p. 44. Polemoniaceae, p. 8. Polemonium eximium, p. 111. Polygonaceae, p. 170. Polygonum bistortoides, p. 110. Polygonum lapathifolium, p. 98. Polygonum minimum, p. 111. Polygonum pensylvonicum, p. 98. Polygonum viviparum, p. 112. Pooecetes gramineus, p. 79. Populus tremuloides, p. 86. Populus trichocarpa, p. 86. Potentilla heanii, p. 87. Primula capillaris, p. 87. Primula parryi, p. 86, 93. Progne subis, p. 78. Prosopium abyssicola, p. 23. Prosopium gemmiferum, p. 23. Prosopium prolixus, p. 23. Prosopium spilonotus, p. 23. Prosopium williamsoni, p. 17, 23, 27, 36. Prunus emarginata, p. 86. Pseudoacris triseriata, p. 49. Pseudtsuga menziesii, p. 11, 100, 148. Pterostegeal, p. 186. Pterostcgia, p. 170. Ptychocheilus lucius, p. 23. Ptychochcilus oregonensis, p. 23, 33. Purshia tridentata, p. 86. Pyrocephalus rubinus, p. 140. ()uercu.s gambelii, p. 11, 86, 98, 100. ()j/erc?/.s' turbinella, p. 11. fian<7 pretiosa, p. 49, 51. Ranunculus, p. 11. Ranunculus jovis, p. 93. Rcgulus calendula, p. 79, 148. Regulus satrapa, p. 79, 148. Rclictus solitaries, p. 24, 27. Rhinichthys cataractac, p. 17, 24, 27. Rhinichthys falcatus, p. 24. Rhinichthys oscidus, p. 17, 24, 27, 28, 36. Rhinichthys sp. p. 24. Rhniocheilus lecontei, p. 53. fi//K'.v cereum, p. 86. ftj'/w wolfii, p. 86. Richardsonius balteatus, p. 17, 23, 27, 36. Richardsonius cgrcgius, p. 23, 36. Rorripa, p. 8. Rubus pair i floras, p. 94. Salicomia, p. 161. Salix arctica, p. 110, 112. Salmo apache, p. 22. SaZmo cZarfci, p. 17, 22, 27, 28, 36. Salmo gairdneri, p. 22. Salmo sp., p. 22. Sahnonidas, p. 22. Salpinctes obsoletus, p. 79. Salsola, p. 161. Salvadora hexdepis, p. 48. Salvelinus malma, p. 22. Sarcobatus, p. 161. Saxifraga, p. 11. Saxifraga cespitosa, p. 112, 115. Saxifraga flagellaris, p. 112. Sayornis nigricans, p. 140. Sclcroporus elongatus, p. 51. Sclcroporus graciosus, p. 50. Sclcroporus magister, p. 47. Sedum rosea, p. 111. Selaginella selaginoides, p. 87. Selasphorus platycercus, p. 78, 157. Senedo midtilobatus, p. 10. Senecio streptanthifolius, p. 10. Sequoiadendron giganteum, p. 86. SiaZta currucoides, p. 79, 150. Sifl/io mexieana, p. 79, 140, 150. Sibbaldia procumbens, p. 110. Silene clokcyi, p. 87. Sitanion hystrix, p. 86. Sitta canadensis, p. 78, 148. Sitta carolinensis, p. 78. Sifta pygmaea, p. 78. Sonom semianulata, p. 48, 49. Sorex /[/c//i, p. 225. Sorex palustris, p. 221, 225. Somv tenellus, p. 221, 225. Sorex trowbridgei, p. 225. Sorex vagrans, p. 221, 225. Spermophilus armatus, p. 225. Spermophilus beldingi, p. 221, 225. Spermophilus lateralis, p. 221, 225. Sphyrapicus ruber, p. 144, 157. 268 GREAT BASIN NATURALIST MEMOIRS No. 2 Sphyrapicus thyroideus, p. 78, 147. Sphyrapicus varius, p. 78, 147, 157. Spinus pinus, p. 148. Spizella passerina, p. 80. Stanleya, p. 8. Stellula calliope, p. 78. Stenogonum, p. 170. Stetophaga ruticilla, p. 140. Stipa californica, p. 86. Streptanthus, p. 8. Sfn'x occidentalis, p. 77. Suaeda, p. 161. Sylvilagus nuttallii, p. 221, 225. Synthyris ranunculina, p. 87. Tachycineta bicolor, p. 78. Tachycineta thalassina, p. 78. Tamiasciurus douglasii, p. 225. Tamiasciurus hudsonicus, p. 225. Tanacetum compaction, p. 87. Taxus brevifolia, p. 86. Tegeticula, p. 3. Thamnophis cyrtopsis, p. 48, 49. Thalictntm fendleri, p. 86. Thamnophis elegans, p. 49. Thelypodium, p. 8. Thermopsis montana, p. 86. Thryomanes bewickii, p. 140. Toxostoma dorsale, p. 140. Trifolium andersonii, p. 86. Tri folium monoense, p. 110. Trisetum spicatum, p. 110, 112, 115. Troglodytes aedon, p. 78, 140. Tsuga mertensiana, p. 86. Tardus migatorius, p. 79, 140. Tyrannus tyrannus, p. 140. Tyrannus verticalis, p. 140. Urosaurus ornata, p. 47. t/ta stansburiana, p. 47. Valeriana occidentalis, p. 93. Vermivora celata, p. 79, 140, 157. Vermivora luciae, p. 140. Vermivora ruficapilla, p. 157. Vermivora virginiae, p. 79. 157. Viguiera, p. 7. Vio/a adunca, p. 86. Virgo beUu, p. 140. Virgo gilvns, p. 79, 140. Virgo olivaceus, p. 140. Virgo solitarius, p. 79, 157. Wi/sonia pusilla, p. 79, 157. Wyethia, p. 7. Xanthium, p. 85. Xyrauchen texanus, p. 25, 33. Yucca, p. 3. Zapus princeps, p. 225. Zenaida asiatica, p. 140. Zonotrichia iliaca, p. 80. Zonotrichia leucophrys, p. 80. Zonotrichia lincolnii, p. 80. Zonotrichia melodia, p. 66, 67, 80. NOTICE TO CONTRIBUTORS Original manuscripts in English pertaining to the biological natural history of western North America and intended for publication in the Great Basin Naturalist should be di- rected to Brigham Young University, Stephen L. Wood, Editor, Great Basin Naturalist, Provo, Utah 84602. Those intended for the Great Basin Naturalist Memoirs should be sim- ilarly directed, but these manuscripts are not encumbered by a geographical restriction. Manuscripts. Two copies of manuscripts are required. They should be typewritten, double spaced throughout on one side of the paper, with margins of at least one inch on all sides. Use a recent issue of either journal as a format, and the Council of Biology Edi- tors Style Manual, Third Edition (AIBS 1972) in preparing the manuscript. An abstract, about 3 percent as long as the text, but not exceeding 200 words, written in accordance with Biological Abstracts guidelines, should precede the introductory paragraph of each ar- ticle. Authors may recommend one or two reviewers for their article. All manuscripts re- ceive a critical peer review by specialists in the subject area of the manuscript under con- sideration. Manuscripts that are accepted and that are less than 100 printed pages in length will automatically be assigned to the Great Basin Naturalist. Those manuscripts larger than 100 printed pages in length will be considered for the Memoirs series. Illustrations and Tables. All illustrations and tables should be made with a view toward having them appear within the limits of the printed page. Illustrations that form part of an article should accompany the manuscript. Illustrations should be prepared for reduction by the printer to either a single-column (2 5/8 inches) or double-column (5 1/2 inches) width, with the length not exceeding 7 1/2 inches. Costs Borne by Contributor. Contributors to the Great Basin Naturalist should be pre- pared to donate from $10 to $30 per printed page toward publication of their article (in addition to reprint costs outlined in the schedule below). Authors publishing in the Great Basin Naturalist Memoirs may be expected to contribute $35 per printed page in addition to the cost of printed copies they purchase. No printed copies are furnished free of charge. A price list for reprints and an order form are sent with the galley proof to contributors. Reprint Schedule of the Great Basin Naturalist 2 pp. 4 pp. 6 pp. 8 pp. 10 pp. 12 pp. Each 100 copies $20 $24 $28 $32 $36 $40 additional 200 copies 28 32 36 40 44 48 2 pp. 300 copies 36 40 44 48 52 56 $4 Great Basin Naturalist Memoirs No. 1 The birds of Utah. By C. L. Hayward, C. Cottam, A. M. Woodburrv, H. H. Frost. $10. No. 2 Intermountain biogeography: A symposium. By K. T. Harper, J. L. Reveal, et al. $15. tfg^v cflcme Bookbinding Co., Inc. 100 Cambridge St. Charlestown. MA 02129 3 2044 072 231 384 Date Due