Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

A Natural History of California: Second Edition
A Natural History of California: Second Edition
A Natural History of California: Second Edition
Ebook1,500 pages20 hours

A Natural History of California: Second Edition

Rating: 0 out of 5 stars

()

Read preview

About this ebook

In this comprehensive and abundantly illustrated book, Allan A. Schoenherr describes the natural history of California—a state with a greater range of landforms, a greater variety of habitats, and more kinds of plants and animals than any area of equivalent size in all of North America. A Natural History of California focuses on each distinctive region, addressing its climate, rocks, soil, plants, and animals.
 
The second edition of this classic work features updated species names and taxa, new details about parks reclassified by federal and state agencies, new stories about modern human and animal interaction, and a new epilogue on the impacts of climate change. 
LanguageEnglish
Release dateJul 3, 2017
ISBN9780520964556
A Natural History of California: Second Edition
Author

Allan A. Schoenherr

Allan A. Schoenherr is Professor Emeritus at Fullerton College, where for decades he taught courses on ecology, zoology, and the natural history of California. He also taught ecology and natural history courses at the University of California, Irvine; California State University at Fullerton; and the Semester at Sea program. He is the coauthor of Natural History of the Islands of California and Wild and Beautiful: A Natural History of Open Spaces in Orange County and is coeditor of Terrestrial Vegetation of California, Third Edition.

Related to A Natural History of California

Related ebooks

Earth Sciences For You

View More

Related articles

Related categories

Reviews for A Natural History of California

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    A Natural History of California - Allan A. Schoenherr

    A NATURAL HISTORY of CALIFORNIA

    A NATURAL HISTORY of CALIFORNIA

    Second Edition

    Allan A. Schoenherr

    UC Logo

    UNIVERSITY OF CALIFORNIA PRESS

    University of California Press, one of the most distinguished university presses in the United States, enriches lives around the world by advancing scholarship in the humanities, social sciences, and natural sciences. Its activities are supported by the UC Press Foundation and by philanthropic contributions from individuals and institutions. For more information, visit www.ucpress.edu.

    University of California Press

    Oakland, California

    © 2017 by The Regents of the University of California

    Library of Congress Cataloging-in-Publication Data

    Names: Schoenherr, Allan A., author.

    Title: A natural history of California / Allan A. Schoenherr.

    Description: Second edition. | Oakland, California : University of California Press, [2017] | Includes index.

    Identifiers: LCCN 2017005894 (print) | LCCN 2017008534 (ebook) | ISBN 9780520295117 (cloth : alk. paper) | ISBN 9780520290372 (pbk. : alk. paper) | ISBN 9780520964556 (epub and ePDF)

    Subjects: LCSH: Natural history—California.

    Classification: LCC QH105.C2 S36 2017 (print) | LCC QH105.C2 (ebook) | DDC 508.794—dc23

    LC record available at https://lccn.loc.gov/2017005894

    Manufactured in United States of America

    25  24  23  22  21  20  19  18  17  

    10  9  8  7  6  5  4   3  2  1

    CONTENTS

    Acknowledgments

    Preface

    Introduction

    Plates

    1 • CALIFORNIA’S NATURAL REGIONS

    2 • BASIC ECOLOGY

    3 • BASIC GEOLOGY

    4 • SIERRA NEVADA

    5 • MOUNTAINTOPS

    6 • PACIFIC NORTHWEST

    7 • COAST RANGES

    8 • CISMONTANE SOUTHERN CALIFORNIA: MAINLAND AND ISLANDS

    9 • CALIFORNIA’S DESERTS

    10 • THE GREAT CENTRAL VALLEY

    11 • INLAND WATERS AND ESTUARIES

    Epilogue

    Noteworthy Publications

    Index

    ACKNOWLEDGMENTS

    For fear of forgetting someone, I am tempted not to acknowledge anyone’s help on this book. However, with a project of this size, it is obvious that quite a few people had something to do with the outcome. With this in mind, I hope that the friends, colleagues, and acquaintances that were inadvertently overlooked will understand how much I appreciate their help.

    For starters, I am indebted to the administration of Fullerton College for providing me with a sabbatical leave, during which time I wrote the first draft. Second, Debbie Horrocks and Amy Gigliotti converted the typewritten first draft to word-processed versions. I am also deeply indebted to Fullerton College computer wizards such as Kent Gordon, Co Ho, Vinh Ho, Geoff Clifton, and Bill Dalphy, who tirelessly gave their time to jury-rig equipment and write programs enabling me to incorporate numerous revisions with minimal effort and convert all my computer files to a single format.

    Unless otherwise noted, all photographs are mine. For contributions of art, I appreciate the work of Karlin Grunau Marsh, Phil Lingle, Geoff Smith of Fullerton College, Pat Brame of Eaton Canyon Nature Center, and Philip Brown of the Southwestern Herpetologists Association. Further thanks go to those many persons cited in text who gave me permission to use work from previously published materials.

    For their sage advice and comments on writing style, I am deeply appreciative to the following persons: Diana Cosand from Chaffee College and Chuck Leavell of Fullerton College, for their help with the sections on basic ecology; Peter Tresselt, formerly of Fullerton College, Rick Lozinsky of Fullerton College, and N. King Huber of the US Geological Survey, for their help with the sections on geology; Alan Romspert of the California State University Desert Studies Consortium and Lenny Vincent of Fullerton College, for help with Chapter 9, on deserts; and Phil Pister of the California Department of Fish and Wildlife for his extremely helpful review of Chapter 11, on inland waters. In preparation of the second edition, I am indebted to colleagues such as Mick Bondello, Chuck Leavell, Doug Allan, Rick Lozinsky, and Lenny Vincent who kindly read and made beneficial suggestions.

    Finally, I am indebted to Art Smith and Ernest Callenbach, editors from the University of California Press, who provided invaluable assistance and encouragement at all stages in the development of the first edition. I am thankful for the encouragement of Blake Edgar from the University of California Press, without which I would not have undertaken the effort to do a second edition.

    PREFACE

    In 1992, when the first edition of this book was published, the University of California Press, in celebration of entering the second century of publishing, honored A Natural History of California among 100 Centennial Books published between 1990 and 1995. A special imprint opposite the title page declared that special honor. It has been 25 years since the first edition of this book was written. During that time, several things have changed that should be addressed. For starters, associated with a revolution in DNA technology, there has been a serious realignment of biological taxa. Not only have many species names been changed, but there has also been significant reassignment within families. Several strange relationships have appeared. For example, Maples and Buckeyes are now in the Soapberry family, Sapindaceae. Some familiar families have been broken up. For example, the Figwort or Snapdragon family, Scrophulariaceae, is now broken up into three families. Bee Plant, Scrophularia, and Mullein, Verbascum, are still Scrophs, but the Bush Monkeyflowers, Mimulus spp., are in the Lopseed family, Phrymaceae. The Broomrape family, Orobanchaceae, now includes Broomrape (Orobanche), Bird’s Beak (Cordylanthus), Lousewort (Pedicularis), and Paintbrush and Owl’s-clover (Castilleja spp.). The Plantain family, Plantaginaceae, includes Snapdragons (Antirrhinum), Chinese-houses (Collinsia), Ghost Flower (Mohavea), Bush Penstemon (Keckiella), and all the herbaceous Penstemons (Penstemon spp.). The Lily family, Liliaceae, has been split into at least six families. Remember the Lily family? About all that remains in the Lily family are Mariposa Lilies (Calochortus spp.). Now, Agaves, Yuccas, and Desert Lilies (Hesperocallis) are in the Agavaceae. Onions and garlics are in the Alliaceae. Beargrasses (Nolina spp.) are in the Butcher’s-broom family, Ruscaceae. The Brodiaea family, Themidaceae, includes Brodiaeas (Brodiaea spp.), Goldenstars (Bloomeria spp.), and Blue Dicks (Dichelostemma spp.). Most of the garden varieties such as Daffodils, Paper Whites, Narcissus, and Naked Ladies are now in the Amaryllis family, Amaryllidaceae.

    There have been many changes in animal taxonomy as well. Not only have many of the familiar genera and species names been changed, but there have been family realignments as well. For example, in the past, many of our familiar lizards have been in the Iguana family, Iguanidae. Desert Iguanas (Dipsosaurus) and Chuckwallas (Sauromalus) are still Iguanas, but the Collared Lizards (Crotaphytus) and the Leopard Lizards (Gambelia) are now in the Crotaphytidae. All the rest, at least seven genera of common lizards, including Fence Lizards (Sceloporus) and Side-blotched Lizards (Uta), are now in the Horned Lizard family, Phrynosomatidae.

    Essentially all of the traditional field guides are now out of date. In order to compensate for that problem, I often will include the new scientific name followed, in parentheses, by the former scientific name of the various plants and animals that I discuss. The concept, for naming purposes, of making scientific names permanent by using the dead languages Greek and Latin, in recent years, has been upended with so many name changes. It is a bit ironic that the unofficial common names of many species have become more permanent than the scientific names.

    Another big change for California is that many of the regions and parks that I discuss have been reclassified by federal and state agencies. For example, Death Valley, Joshua Tree, and Pinnacles have been enlarged and upgraded from National Monuments to National Parks, and a large region of the east Mojave Desert has been established as the Mojave National Preserve. Furthermore, a number of lands administered by the US Forest Service and the Bureau of Land Management (BLM) have been reclassified. For example, the San Gabriel Mountains, Berryessa Snow Mountain, and Giant Sequoia in the Sierra National Forest have been upgraded to National Monuments and will be administered by the National Park Service.

    In 1994, with passage of the Desert Protection Act, the total area of California classified as wilderness was brought to 14 million acres (56.656 km²). In addition to the national parks mentioned above, 69 Wilderness Study Areas on BLM land, many of them scenic isolated mountain ranges, officially became classified as wilderness. In February 2016, three new national monuments in the Mojave Desert added 1.8 million acres (7284 km²) of federal protection.

    The specter of climate change also has changed much of the information about California’s natural areas. Drought and warmth have stressed our ecosystems to the point that the future of California and its landscapes must be discussed. Fire regimes and intensity have changed. Species composition of ecological communities has changed, and will continue to change. Uphill and northward distribution of plant communities is already underway. Invasion of nonnative species has altered our natural plant communities, which also changes species composition and fire frequency. The epilogue addresses these issues.

    Although not necessarily new, there have also been changes in ecological thinking. As the distribution and composition of natural communities becomes changed, there is a group of ecological concepts that need to be addressed. Subjects such as ecological islands, patch size, habitat fragmentation, keystone species, mesopredator release, and trophic cascades will be incorporated where appropriate.

    Finally, research on animals, plants, and their interactions in ecosystems has continued during the interval since publication of the first edition of A Natural History of California. There are many new stories to tell about natural history in California and those stories will be incorporated as much as possible in this new edition. Hopefully readers will find this material as useful and interesting as they have in the past.

    INTRODUCTION

    A natural history is an account of natural phenomena. Over 300 years before Christ was born, Aristotle wrote his Historia Animalium, a series of nine books on the anatomy and habits of many animals native to Greece. In A.D. 77, Pliny the Elder, a Roman, wrote the 37-volume Natural History in an attempt to compile an encyclopedia of all known natural phenomena. For over 1500 years, this work served as the basic reference for information about nature. Ever since then, the expression natural history has been used to refer to a description of living organisms, their habits, and how they relate to the environment.

    A more modern term used to describe organisms and their relationships with the environment is ecology. The literal meaning of this word is study of the house. In this context, house’ means the environment. This book easily could have been entitled The Ecology of California, because it is about the creatures and the environment in the state of California. In recent years, however, ecology has become a highly theoretical and technical science, and it does not encompass a study of rocks and geologic history. This book includes a description of organisms, rocks, and the environment, and an attempt to explain what factors through time created what we see today. The facts behind the concepts that are stressed in this book have been drawn from ecology, zoology, botany, biogeography, climatology, paleontology, and geology. Therefore, it seems best to call it a natural history.

    California is a highly complex geographic unit. There is more climatic and topographic variation in California than in any other region of comparable size in the United States. The highest and lowest points in the lower 48 states are less than 100 mi (160 km) from each other in eastern California. Mount Whitney, figure 4.11 located on the crest of the Sierra Nevada, at an elevation of 14,505 ft (4421 m), is only 84.6 mi (136.2 km) west of Badwater in Death Valley, which is 282 ft (86 m) below sea level. Interestingly, the US Geological Survey brass benchmark on the summit of Mount Whitney indicates an elevation of 14,494 ft (4418 m). The new designation is the product of new techniques for estimating elevation.

    California also has a great range of climates. The Sunset Western Garden Book (Lane Publishing Co.) describes 24 different climatic zones within the state. Outside of California, few states have more than four zones, and most have only one.

    Total precipitation per year of more than 120 in (300 cm) may commonly occur in the northwestern forests. Honeydew in northern California has recorded the highest average annual precipitation of 104.18 in (264.6 cm), although the official record for annual precipitation, 161 in (403 cm), occurred in the Santa Lucia Mountains of the southern Coast Ranges. In the southwestern deserts, it is not uncommon for some locations to go several years without measurable precipitation. Bagdad, a now-deserted community on old Route 66 in the Mojave Desert, was reported to be the driest place in the United States. No measurable precipitation was recorded for 25 months. The lowest official record for annual precipitation of 1.6 in (4.06 cm) is held by Death Valley.

    The range of temperatures in California is also extreme. Subzero temperatures may continue for many days at high elevations in the mountains, and Furnace Creek in Death Valley, at 134°F (57°C), holds the record for the highest official air temperature ever recorded. A slightly higher temperature for a location in Libya is sometimes touted as the highest, but in recent years it has been determined that the record was erroneous.

    The geologic picture of California is also very complex. Most of the state is composed of a mélange of rock units from different sources, and of different ages, that became attached to its western border as North America slid westward over the floor of the Pacific Ocean. Tremendous forces have stretched and warped the land, so that California today is a mixture of mountains and valleys of diverse origin cut through by major fault systems. Lands west of the San Andreas fault, for example, have slid to their present position from a point adjacent to mainland Mexico, perhaps as much as 300 mi (480 km) to the south. The dominant topographic feature of California, the Sierra Nevada, which in eastern California runs on a predominantly north to south direction for approximately 400 mi (640 km) is composed primarily of granitic rocks. The great variety of rock materials degrade to form a corresponding variety of soils. Soils of different mineral content and texture have a profound influence on plants and animals. A particular soil type may have its own specially adapted community of organisms.

    The combination of diverse climate and diverse soil is responsible for the development of a diversity of habitats. The total number of habitat types varies, depending on whether the classifier is a splitter or a lumper. For example, the California Natural Heritage Section recognizes about 300 natural communities, and the California Department of Fish and Wildlife recognizes 178 major habitat types. This book shall consider about 35 terrestrial communities modified from a system developed by Munz and Keck in 1959. Furthermore, about 15 different aquatic communities will be discussed.

    Many specialized plants and animals are found in California. There are more endemic species in California than in any area of equivalent size in North America. There are more than 6000 native plants in California, and about a third of them occur naturally nowhere else on earth. Among vertebrate animals, there are nearly 1000 native species. Depending on how a person classifies the species, that number includes about 540 birds, 214 mammals, 77 reptiles, 47 amphibians, and 83 freshwater fishes. Of these, 65 species are considered to be endemic. Among animals as a whole, including insects and other invertebrates, at least 50% of the species and subspecies are confined within the borders of the state.

    The reason for writing a book of this type is to familiarize readers with this special place called California. After reading it, a person should be able to describe the climate, rocks, soil, plants, animals, and biogeography of any area in California. The reader should be able to explain how those things got there and the ways in which they relate to each other.

    Another, no less important reason for writing a book of this type is to foster appreciation for California’s natural diversity. Much of California’s nature is threatened. The Nature Conservancy has reported that about 25% of the state can no longer support its original communities of plants and animals. Since 1900, 65% of the state’s coastal wetlands have been dredged, filled, or drained. Riparian forests of the Great Central Valley, which once covered hundreds of square miles, are nearly gone, and some specialized communities, such as Southern Coastal Dune Scrub, have been reduced to only a few hundred acres. Not only does California have the greatest number of unique organisms in the continental United States, but it also has the highest number of threatened species, many of which are among the unique or endemic ones. Over 600 kinds of plants are listed as threatened with extinction. The greatest threat to California’s natural ecosystems is growth of the human population. California already is the most populous state. About 70% of these people live within a few miles of the coast, with the greatest concentration of humans in the San Francisco Bay area and South Coast from Los Angeles to San Diego. The Great Central Valley, while mostly associated with agriculture, also has population centers such as Fresno and Sacramento. So, in spite of the threats mentioned above, over half of California’s landscape is comparatively undisturbed. According to the California Protected Areas Database, 52% is public land, and 46.7% of California is classified as protected. California has more officially designated wilderness, 14.36% of its area, than any state outside of Alaska.

    This book is organized around geographic regions using, as a starting point, 12 geomorphic provinces described by the California Geological Survey (previously known as California Division of Mines and Geology). Superimposed on these regions are the natural biotic provinces based on climate and living organisms. The result is a sequence of chapters in which various regions of the state are characterized with respect to climate, geology, and biotic communities. Animals and plants are discussed using recent standardized common and scientific names wherever possible. Standardized vocabulary is also used for geologic formations.

    In no part of this book is the treatment of plants, animals, and geologic units meant to be exhaustive. In each region, the emphasis is on those things that are conspicuous, distinctive, and/or interesting. In cases where an organism, a community, or a rock unit occurs in more than one geographic area, it is discussed in the context in which it seems to be most significant.

    Every attempt has been made to ensure that facts and concepts presented here are based on the most up-to-date references. However, there is a good deal of disagreement among experts, particularly about evolutionary relationships, biogeography, and geologic history. If errors or discrepancies seem to occur, or if more information is desired, the reader is encouraged to consult the references at the back of the book.

    PLATE 1 California’s generalized vegetation cover (from Keeler-Wolf, T. 2003. An Atlas of the Biodiversity of California . State of California, The Resources Agency, Department of Fish and Game, p.19).

    PLATE 2 Distribution of selected bedrock types (from Kauffman, E. 2003. An Atlas of the Biodiversity of California. State of California, the Resources Agency, Department of Fish and Game, p.17).

    PLATE 3A Yosemite National Park, Mixed Coniferous Forest in winter.

    PLATE 3B Emerald Bay State Park, Lake Tahoe

    PLATE 4A Joshua Tree Woodland and granitic boulders, Joshua Tree National Park.

    PLATE 4B Cactus Scrub featuring Ocotillo, Barrel Cacti, and Desert Brittlebush, Anza-Borrego Desert State Park.

    PLATE 5 General Sherman, the largest tree in the world, in winter, Sequoia National Park.

    PLATE 6A Mixed Coniferous Forest featuring Jeffrey Pine and Wilson Butte, a volcanic dome, north of Mammoth Lakes.

    PLATE 6B Montane Forest featuring Lodgepole Pines and Mountain Hemlock to left of center against horizon, McCabe Lake, Yosemite National Park.

    PLATE 7A Subalpine Forest featuring Foxtail Pines, Sequoia National Park

    PLATE 7B Old Bristlecone Pine in the Patriarch Grove, White Mountains.

    PLATE 8A Alpine Meadow in Humphreys Basin, Inyo National Forest, featuring lupines in bloom and Mount Humphreys in background.

    PLATE 8B Autumn color, featuring Quaking Aspens in a Riparian Woodland, North Lake on Bishop Creek.

    PLATE 9A Dry Mixed Coniferous Forest dominated by Jeffrey Pines and Manzanitas. Mount Lassen as viewed from the northeast.

    PLATE 9B Old-growth Douglas fir, Siskiyou Mountains.

    PLATE 9C Moist subalpine community of Salmon Mountains at Cuddihy Lakes.

    PLATE 10A King Range at the Lost Coast north of Shelter Cove, Humboldt County.

    PLATE 10B Mixed Evergreen Forest, northern Coast Ranges.

    PLATE 11A Eel River and Coast Redwoods in Humboldt Redwoods State Park.

    PLATE 11B Oak Woodland featuring Blue Oaks and Gray Pines, Del Valle Regional Park.

    PLATE 11C California Poppies, Eschscholzia californica, the California State Flower in the Antelope Valley California Poppy Reserve, Los Angeles County.

    PLATE 12A Coastline at Big Sur.

    PLATE 12B Julia Pfeiffer Burns State Park. McWay Falls drops directly into the ocean.

    PLATE 12C Bristlecone Fir, Abies bracteata , endemic to the Santa Lucia Mountains in Monterey County.

    PLATE 13A Coastal Sage Scrub in autumn, showing diversity of species represented by various colors.

    PLATE 13B Oak Woodland on Santa Catalina Island featuring tree-sized Island Scrub Oaks, Quercus pacifica.

    PLATE 14A Mixed Coniferous Forest on the north side of the San Bernardino Mountains featuring Sugar Pines and White Firs.

    PLATE 14B Wildflowers after a fire.

    PLATE 15A Great Basin Sagebrush in Owens Valley. Sierra Nevada in background.

    PLATE 15B Creosote Bush Scrub, the dominant plant community of the warm deserts in California.

    PLATE 16A Cactus Scrub in Anza-Borrego Desert State Park.

    PLATE 16B Thousand Palms Oasis, featuring California Fan Palms, Washingtonia filifera .

    PLATE 17A Vernal pool surrounded by a ring of flowers, endemic Downingia sp.

    PLATE 17B Paternoster of cirque lakes. View southward toward Sawtooth Peak from Black Rock Pass, Sequoia National Park.

    PLATE 17C Thousand Island Lake, a large tarn in a glacial scour. Mount Banner and Mount Ritter with small glaciers in background.

    PLATE 18A California Ground Squirrel, Otospermophilus beecheyi .

    PLATE 18B Chickaree or Douglas’s Squirrel, Tamaisciurus douglasii.

    PLATE 18C Chickaree or Douglas’s Squirrel, Tamiasciurus douglasii .

    PLATE 18D Golden-mantled Ground Squirrel, Callospermophilus lateralis .

    PLATE 18E White-tailed Antelope Squirrel, Ammospermophilus leucurus .

    PLATE 18F Yellow-pine Chipmunk , Tamias amoenus .

    PLATE 18G Belding’s Ground Squirrel, Urocitellus beldingi.

    PLATE 18H Desert Woodrat, Neotoma lepida . (photograph by Greg Stewart).

    PLATE 19A California Quail, Callipepla californica .

    PLATE 19B Hooded Oriole, Icterus cucullatus .

    PLATE 19C Wood Duck, Aix sponsa .

    PLATE 19D Western Bluebird, Sialia mexicana .

    PLATE 19E Red-winged Blackbird , Agelaius phoeniceus .

    PLATE 19F Western Scrub-jay, Aphelocoma californica .

    PLATE 19G Western Tanager, Piranga ludoviciana .

    PLATE 19H Yellow-billed Magpie, Pica nuttalli .

    PLATE 20A Paintbrush, Castilleja sp.

    PLATE 20B Scarlet Gilia, Ipomopsis aggregata .

    PLATE 20C Harlequin Lupine, Lupinus stiversii .

    PLATE 20D Leopard Lily, Lilium pardalinum .

    PLATE 20E Red Columbine, Aquilegia formosa .

    PLATE 20F Alpine Columbine, Aquilegia pubescens .

    PLATE 20G Broad-leaved Helleborine, Epipactis helleborine .

    PLATE 20H Purple Owl’s Clover, Castilleja exserta.

    1

    California’s Natural Regions

    THE CALIFORNIA GEOLOGICAL SURVEY (previously known as the California Division of Mines and Geology) has divided the state into 12 geomorphic provinces based on rock type and topography ( figure 1.1 ). The words topography and geomorphic refer to the shape of the land: topography means place picture, and geomorphic means earth form. These geomorphic provinces represent natural units within which the boundaries of landforms are remarkably consistent with those of biological communities. That is, the shape of the earth influences climate, and climate influences the distribution of plants and animals. I have adapted geomorphic provinces, with some modifications, into natural regions which shall provide the framework of organization for this book ( figure 1.2 ).

    FIGURE 1.1 Geomorphic provinces (from Hill 1984).

    FIGURE 1.2 Landforms and natural regions of California (courtesy of California Insect Survey, Department of Entomology and Parasitology, University of California).

    Reference to landforms is way to refer to geographic features. They can be illustrated by means of aerial photographs or they may be depicted in the form of a relief map that shows topography, highlighting mountains and valleys (figure 1.2). Outlines of the landforms on a relief map show how California can be divided into natural regions which will be the subject of most of the chapters in this book. Other authors have divided the state into natural regions using different criteria. Some of these divide the state into fewer regions and others include more categories. Peter Berg and Raymond Dasmann in the 1970s divided the state into bioregions, which roughly correspond to the natural regions to which I refer (Table 1.1). M.D.F. Udvardy in 1975 described biogeographical provinces of the world, and Robert Bailey in 1976 introduced the idea of ecoregions. In 1983, he defined 19 ecoregions in California. The concept behind bioregions or ecoregions is that ecological boundaries need not correspond to political boundaries. Bioregions are supposed to be based on a series of criteria including humans living in harmony with the natural environment. W.J. Barry in 1991 described 24 ecological regions and Hartwell Welsh, who was associated with the US Forest Service, in 1994, divided the state into 16 bioregions, while other authors refer to 10 or 6. Adoption of the bioregion system, therefore, has been inconsistent. Some authors would divide the coastal mountains into three parts, north, central, and south, using the term South Coast to refer to the region I call Cismontane Southern California, which includes the Transverse and Peninsular Ranges. Other authors would split the Coast Ranges in the San Francisco area with a region called Bay/Delta. I recognize three desert regions, Great Basin, Mojave, and Colorado, from north to south. Others do not recognize the Great Basin Desert in California, instead referring to the areas as Eastern Sierra and/or Modoc in reference to the Modoc Plateau in the northeastern corner of the state. Most authorities refer to the southernmost desert as the Colorado Desert, which is the California part of a larger desert unit, the Sonoran Desert, which also occurs in southern Arizona, Baja California, and the state of Sonora in northwestern Mexico. Some would not recognize the Colorado Desert at all, but instead refer simply to the Sonoran Desert in California. Bailey divided the Colorado Desert of California into Sonoran and Colorado subdivisions.

    TABLE 1.1

    A Comparison of Natural Regions with Bioregions

    CLASSIFICATIONS OF BIOTIC COMMUNITIES

    Dividing the state into regions based on climate and geography leads to classifying its biotic communities. It has been recognized for many years that different vegetation is associated with different types of climate. Various classifications of vegetation, based on climatic data, have pervaded scientific and lay literature. Table 1.2 compares three such systems that apply to California’s flora.

    TABLE 1.2

    A Comparison of Three Systems of Community Classification

    One of the earliest attempts was the life zone system of C. Hart Merriam. This system, published in 1892, was based primarily on temperature. The zones were named for geographic areas that had temperature regimes similar to that which Merriam observed in Arizona as he traveled from the floor of the Grand Canyon to the summit of nearby San Francisco Peaks. Merriam worked for the US government, and his system became entrenched in government communications. It is still used by many naturalists who are employed by various federal and state agencies. The correlation between elevation and latitude has merit, but the fact that the system is based primarily on temperature and infers precipitation makes it truly applicable only to the southwestern United States.

    Two ecologists, Clements and Shelford, popularized the biome concept. Biomes are large ecosystems in which variations in temperature and precipitation have created a characteristic assemblage of plants and animals. On a broad scale, this system is appropriate, and its vocabulary is in widespread use.

    In 1947, L.R. Holdridge proposed a scheme for classifying world plant formations based on evapotranspiration ratios. It is an eloquent system that applies raw climatic data more precisely than the biome system. The Holdridge scheme, however, lacks simplicity and has not gained widespread use.

    California has been named one of the world’s top 25 biodiverse regions of the world. In California, the range of climate is so extreme and the vegetation so diverse that broad classification schemes such as world plant formations or biomes have proved inadequate. Instead, the system of plant communities, published by Phillip Munz and David Keck in 1959, became a system of choice used by many ecologists in California. This system is based on one or more dominant plant species that occupy each area. It is broad enough that there are not too many categories to remember, and it is precise enough that its categories encompass meaningful units. Originally, the system included only 29 plant communities. Throughout this book, the Munz and Keck system of plant communities, with some modification to include about 35 categories, will provide a basic frame of reference.

    Robert Bailey in 1983 developed a system of ecoregions that is used today by the US Forest Service. He divided California into 19 ecoregions. Using categories that are similar to ecoregions, the Jepson Manual of the Vascular Plants of California divides the state into 50 geographic subdivisions. These categories are arranged hierarchically into Provinces, Regions, and Subregions. The categories are defined on the basis of geography, climate, and vegetation. The manual uses these 50 subdivisions to describe the distribution of California plants. These categories are a bit too refined for use in this book, but they will not be ignored entirely.

    According to the Jepson Manual, there are three floristic provinces: California, Great Basin, and Desert. The California Floristic Province is the primary group of plants in the state, and is distinctly Californian. It includes the most diverse group of plants in the state. The province essentially refers to everything west of the deserts, and it encompasses that part of California which is profoundly influenced by our Mediterranean climate, characterized by winter precipitation and dry summers.

    A chain of mountain ranges extending the length of California forms a significant block to winter storms off the Pacific Ocean. From north to south, these ranges are the Cascade Mountains, Sierra Nevada, Transverse Ranges, and Peninsular Ranges. Deserts lie to the east of these ranges. The Great Basin Province is the western edge of the largest desert in North America, which extends eastward across Utah and Nevada all the way to the Rocky Mountains. This is a cold desert in which winter precipitation often arrives as snow. In general, the dominant vegetation is composed of cold-tolerant shrubs such as various forms of sagebrush (Artemisia sp.). The Desert Province is subdivided into the Mojave and Sonoran (Colorado) Deserts. These are warm deserts characterized predominantly by drought-tolerant shrubs such as Creosote Bush (Larrea tridentata). The Mojave gets primarily winter precipitation with snow at higher elevations. California’s Colorado Desert has two rainy seasons with approximately half its precipitation arriving as summer thunderstorms.

    People who classify things could be splitters or lumpers. Lumpers are folks who tend to categorize things on the basis of similarities. Splitters like to see differences between things, and tend to assign a name to each of the categories. Phillip Munz and David Keck were lumpers. In recent years, splitters have developed various systems of plant community classification which may be referred to as vegetation types. In association with the University of California’s Natural Reserve System and the California Department of Fish and Game, Daniel Cheatham and J.R. Haller in 1975 developed a list of habitat types that included about 145 communities. In 1986, Robert Holland working on the California Natural Diversity Database for the California Department of Fish and Game developed a list of 260 terrestrial natural communities. In 1995, the California Native Plant Society published A Manual of California Vegetation by John Sawyer and Todd Keeler-Wolf. This manual was designed as an evolving system that would recognize rare or unrecognized vegetation types as well as refining categories into more specific plant associations and series. Originally, the manual included about 250 vegetation types. In the second edition in 2009, in collaboration with Julie Evans, the number of units called alliances was upped to nearly 500. Many hours of scientific surveys, mapping, and analysis went into preparation of these publications, and they are meant to be the basic reference for descriptions of California vegetation. Unfortunately, identification of many of these alliances in the field requires a certain amount of botanical knowledge, and alliances often are composed of a single species, which makes using the alliance system unwieldy for a book of this type. For example, the common forest community known as Mixed Coniferous Forest (formerly Yellow Pine Forest) is composed of about six common species of trees, each of which may grow in clumps. In the Manual of California Vegetation, this community is split into 10 alliances, basically associated with clumping of individual species. Therefore, without ignoring the needs of botanists, but describing vegetation that is understandable to a lay person, in this edition of the book, I will continue to use a system of recognizable plant communities (with some reference to alliances) that has been used by ecologists and nature writers for many years. To enhance understanding of California’s diverse vegetation, a color-coded map of generalized vegetation cover is depicted in plate 1.

    CLIMATE AND WEATHER

    California’s Mediterranean climate primarily is responsible for its great diversity of biotic communities, although other factors such as soil type and slope exposure are also involved. Unique among climates of the world, in a Mediterranean climate, precipitation occurs primarily during winter months, the coldest time of year and the time with short daylight hours. Since photosynthesis in green plants provides the nutrient basis for food chains and ecosystems, it should be obvious that winter is generally not the optimum time to promote plant growth. Thus, most ecosystems associated with a Mediterranean climate must be considered relatively food-poor, and that dictates specialized adaptations among the plants and animals that inhabit the area.

    The study of climate and weather is known as meteorology. Climate refers to the overall combination of temperature, precipitation, winds, and so forth that any region may experience. Weather refers to daily variations in these phenomena. Changes in temperature and precipitation are functions of many variables. As the earth rotates, light strikes its surface at different angles in different latitudes, and because the earth is tilted on its axis by 24° these angles change with the seasons. During summer in the northern hemisphere, the North Pole is tilted toward the sun so the sun’s rays hit the surface more directly, and the photoperiod (day length) is longer. These differences cause uneven heating, which is expressed as wind. Winds carry water vapor, the amount of which is influenced by differing rates of evaporation and precipitation, depending on the temperature. Furthermore, wind circulation patterns are influenced by the earth’s rotation on its axis and differences in local topography. All this translates into major generalizations about what causes climate and weather for a specific geographic region.

    Major Movement of Air Masses

    Near the equator, sunlight hits the earth directly. On the first day of spring in the northern hemisphere, or the first day of autumn in the southern hemisphere, the sun is directly over the equator, at which time daylight and darkness are 12 hr long all over the world. Wherever the sun shines straight down, there will be a greater degree of warming than anywhere else. Air will rise at that point. Meteorologists indicate that for every 1°F (1.7°C) increase in temperature the atmosphere can hold 4% more moisture. Air on the surface of the earth will move as winds toward the point of rising air from the north and south. As the air rises, it becomes less dense and thus cooler. This is called adiabatic cooling. The precise amount of cooling varies with the amount of water vapor in the air, but on the average it is about 3–5°F/1000 ft (1°C/100 m). A drop in temperature of 20°F (11°C) will decrease the amount of water held by the air by one-half. When the air is cooled, water vapor condenses and falls in the form of rain. As a result, it rains nearly every day, usually in the afternoon, in a belt around the earth at the equator. As summer progresses in the northern hemisphere, the photoperiod gradually increases until the sun is shining directly on the Tropic of Cancer, 24° north of the equator. On the same day the sun is shining on the Tropic of Capricorn in the southern hemisphere, it shines down at its greatest angle. On this day, June 22, it is the summer solstice in the northern hemisphere and the winter solstice in the southern hemisphere: the longest day of the year in the north and the shortest day of the year in the south. The climate in this region between the tropics is therefore said to be tropical.

    Where air descends to the earth, the process reverses. Descending air becomes compressed and heats at about the same rate as it cools when rising. This process causes the air to absorb water vapor. Descending air therefore tends to be dry and causes liquid water to evaporate from the environment. This dry air descends in two belts around the earth, approximately 30° north and south of the equator, just outside the tropics (figure 1.3). A quick glance at a world map will show that deserts lie along this belt, and along the southwestern coasts of each continent lie regions with a Mediterranean climate. When it is summer in the northern hemisphere, this belt of dry, descending air is pushed farther north. It therefore seldom rains in southern California in the summer. Los Angeles lies at about 35° north latitude. Meanwhile, the belt of precipitation has also moved farther north of the equator, creating regions of summer rainfall in the vicinity of the Tropic of Cancer. A climate marked by periods of summer precipitation is called a monsoonal climate. That includes southern Florida in the United States.

    FIGURE 1.3 Simplified model of major movement of air masses (after Barry, R. G., and R. J. Charles. 1982. Atmosphere , Weather , and Climate. London: Methuen. Illustration from Hensson and Usner 1993, p.34).

    It is not really accurate to use the terms warm and cool with respect to air masses. What actually causes air to rise or descend is a change in density. Warm air is less dense, so it will rise. Cool air is comparatively dense, so it will sink. As it sinks and becomes warmed by compression, it gets even denser. Even though it has become warmed, it will continue to descend. It is best, therefore, to compare air masses in terms of density. Dense air pushes downward with greater pressure; thus, it is said to represent a region of high pressure. Conversely, rising air represents a low-pressure region. Meteorologists explain changes in weather as changes in pressure, which are measured by a barometer. They translate predicted changes of pressure into a weather forecast. A reduction in air pressure might imply that a storm, or a mass of rising air, is approaching. Satellite photos show low-pressure areas covered by clouds and high-pressure areas are clear.

    The Influence of Ocean Currents

    The influence of the ocean on California’s climate is profound. California’s coastline is about 1100 mi (1760 km) long. This great range of latitude causes the climate to vary considerably along the coast from north to south. From Crescent City to San Diego, the range of average annual precipitation is from about 80 in (200 cm) to 10 in (25 cm). On the other hand, the proximity of the ocean means that air temperatures are not extreme. At the coast, freezing is rare, and the temperature rarely exceeds 100°F (38°C). At San Francisco, September temperature averages 62°F (17°C), and January temperature averages 51°F (11°C).

    Ocean water along the California coast is generally cold. Off the northern California coast, winter water temperature is usually around 50°F (10°C). In its journey southward, it warms about 7–8°F (3–4°C). During summer, it is only 8–10°F (5–6°C) warmer. Cold ocean water causes frequent fog, and relative humidity remains high most of the time. Plants water themselves with fog drip, and evaporation rates are low. Along the coast, plant communities range from moist Mixed Evergreen Forest in the north to dry Coastal Sage Scrub in the south.

    Ocean water swirls clockwise in the northern hemisphere and counterclockwise in the southern hemisphere (figure 1.4). This swirling, a product of the earth’s rotation, is known as the Coriolis effect. The earth rotates in an easterly direction. That is why the sun appears to move westward, rising in the east and setting in the west. The effect of this eastward motion is that the water at the equator appears to flow westward. As the water moves westward relative to the earth’s surface, it begins to turn toward the right in the northern hemisphere and toward the left in the southern hemisphere. This phenomenon occurs because the earth’s motion is fastest at the equator, where the circumference of the earth is greatest. To make one complete rotation, the land at the equator has to travel farther than at any other place on earth. Because the surface of the earth moves more slowly north or south of the equator, the increased drag makes flowing water turn to the right, clockwise, in the northern hemisphere and to the left, counterclockwise, in the southern hemisphere.

    FIGURE 1.4 Pacific Ocean currents (from Caughman, M., and J. S. Ginsberg. 1987. California Coastal Resource Guide . Berkeley: University of California Press).

    As the water moves westward at the equator, it picks up heat in the region of direct sunlight. Currents of water that come from the equator are comparatively warm, and they flow northward along the eastern edge of a continent in the northern hemisphere. As this water flows through the arctic region, it cools considerably. When the water flows southward again, toward the equator, it flows along the western edge of a continent. A cold current of this type flows southward along the coast of California.

    In the southern hemisphere, ocean water swirls in an opposite direction. Warm currents from the equator flow southward on the eastern edge of a continent. Cold currents from the Antarctic flow northward along the western edge of continents. One bit of consistency to emerge from all this is that water off the east coast of any continent is comparatively warm regardless of the hemisphere and water off the west coast of any continent is comparatively cold.

    A climate dominated by the influence of the ocean is known as a maritime climate. On a daily basis, water temperature fluctuates very little. Therefore, air temperature fluctuations on the nearby land are also moderated. When the water offshore is comparatively cold, as it is on the west coast of the United States, fog forms over the water and adjacent land nearly every day, a condition that is accentuated as you go farther north. In southern California, rain is typically confined to the ocean and a short distance inland. For rain to occur farther inland, the land itself must be significantly cooler than the air. When the offshore current is comparatively warm, as it is on the east coast, it rains frequently on the land, and fog seldom forms.

    Air masses tend to follow the ocean currents because they are also under the influence of the Coriolis force. The circles of flow, however, are interrupted by the descending air at latitudes about 30° north and south. Where air is descending, very little wind flows transversely across the surface of the earth. Years ago, seamen in wooden sailing ships would have to row when they got becalmed in this part of the ocean, so they would throw over some of the cargo to lighten the load. So many horses from the old world were floating belly up that this part of the ocean became known as the Horse Latitudes. A comparable area of little wind, known as the Doldrums, occurs at the equator. Here, the air is rising. In the northern hemisphere, air flowing toward the earth at the 30th parallel turns to the right, forming a clockwise flow. The effect is that air south of the 30th parallel tends to flow southwestward, forming what is known as the Trade Winds, so named because sailing ships could use these winds to carry goods across the Atlantic to the new world. North of the 30th parallel, winds tend to flow from west to east. These winds are known as the Prevailing Westerlies.

    Storms develop as low-pressure areas over the ocean, and are highly dependent on the water temperature. Warm air in a low-pressure cell rises, carrying with it evaporated water. Wind on the surface of the earth therefore blows toward a low-pressure cell from all directions. In the northern hemisphere, air moving across the surface to replace the rising air of a low-pressure cell must turn to the right, which gives the mass a counterclockwise spin. This spinning mass is similar to a spinning top. It moves randomly unless other forces give it a direction. Prevailing winds provide this force. Storms generated in the tropics tend to be blown westward by the Trade Winds, and storms generated over the northern part of the ocean tend to be blown eastward by the Prevailing Westerlies. In North America, major winter storm tracks come from the north Pacific and flow southeastward (figure 1.5A). They spin counterclockwise as they travel down the coast, so they continue to pull moisture off the ocean and dump it on the land. As one moves from north to south, the land warms faster than the water because air changes temperature more rapidly. During summer, when the air and land in southern California is considerably drier and warmer than the water, the storm fizzles because water vapor won’t condense (figure 1.5B). During winter, storm after storm potentially makes its way down the coast before moving eastward across continental North America (figure 1.5A). This is basic reason that the northern half of the state gets two-thirds of California’s precipitation. Fewer storms cause precipitation in southern California and they rarely continue as far south as the center of Baja California. A high-pressure cell is like a dome of dense air over the southern part of the state and that deflects the storms northward and eastward. Baja California still has cold water offshore, but the land seldom experiences measurable precipitation. Because condensation in moist air occurs over cold water, the coast of Baja California is frequently foggy. Plants water themselves with fog drip but cannot depend on rainfall to provide adequate water. In Baja California, therefore, the desert occurs on the coastline.

    FIGURE 1.5 Seasonal weather patterns showing locations of winter and summer high- and low-pressure zones.

    (A) Winter weather pattern.

    (B) Summer weather pattern.

    At high elevation, there are two jet streams, a polar jet stream and a subtropical jet stream, which move eastward around 100 mph (160 kph). These jet streams tend to push storms eastward, carrying weather systems across the continental United States. The polar jet stream usually has more influence on California weather, but these masses of flowing air meander back and forth similarly to flowing water in a river. When the polar jet stream meanders southward, it pushes Pacific storms toward southern California and precipitation increases. During summer, this is an important force that brings precipitation to the forests of the Pacific Northwest and the northern Sierra Nevada. The subtropical jet stream also can meander northward and bring tropical (monsoonal) storms to southern California, a condition that is accentuated when the eastern Pacific Ocean is exceptionally warm. This is known as an El Niño condition, a Spanish word that is derived from a change in weather that also influences the coast of South America. Heavy rain would arrive around Christmas time along the coast of Peru and Chile where climate is desertlike. The name is in reference to the birth of the Christ child at that time of year. El Niño conditions brought heavy rains, essentially double of normal amounts to parts of California in 1978, 1983, 1993, 1998, 2016 and 2017.

    In the summer, it seldom rains in southern California (figure 1.6). The sun has moved northward so that it shines straight down on the Tropic of Cancer. That is where the belt of rising air generates tropical storms. Because the rising air is farther north at this time, so is the descending air, and that keeps southern California dry in the summer. Even in the absence of an El Niño, during late August or September occasional tropical storms move into southern California from the south. These storms are usually carried westward from the east Pacific toward Hawaii, but some of them move into the belt of descending air where there is no surface wind to give them a westward push. These storms can spin erratically (like a top) and may move over the land, dumping a considerable amount of water in a short time. These sporadic storms, known in Baja California as chubascos, are able to move northward. Such thunderstorms over the Colorado Desert and Peninsular Ranges may cause intense local flooding. Some 5–10 in (12–25 cm) of rain may fall in a few hours, representing a large portion of the annual share of precipitation. Some of these tropical storms also come all the way from the Gulf of Mexico.

    FIGURE 1.6 Different weather conditions. Arrows indicate the direction and source of different weather conditions and seasons for the outhern part of California (from Bailey 1966).

    The moist forests of the Pacific Northwest contain a variety of relict species that depend on the relatively predictable supply of summer rain from the tail end of Alaskan storms. Summer rain in southern California, however, may not occur every year. Nevertheless, certain plant species, such as different populations of relict cypress (Cupressaceae), to some degree owe their existence to summer rain.

    The center of the North American continent tends to be dry because it is a long way from the ocean, the main source of water. Pacific storms, during summer, are deflected over the Midwest by the southern belt of high pressure. In winter, strong storm systems carry precipitation all the way to the interior, where the moisture may be added to the storms traveling eastward from the Pacific. If the difference in pressures between the two storm fronts is extreme, at the contact point between the warm southern storms and the cold Pacific storms, there may be a belt of tornados. Furthermore, during summer, thunderstorms from the southern storm belt carry northward across the Gulf of Mexico all the way to the plains states. They may carry as far as California’s Colorado Desert. In late August, these storms often reach as far as Arizona, where locally heavy precipitation is common. The Arizona portion of the Sonoran Desert experiences considerable summer precipitation. This rain falls during the hottest time of the year, however, so that evaporation rates are very high and the water does not sink very far into the ground. The area around Tucson is a desert, whereas the area around Los Angeles is not. Both areas receive approximately the same amount of precipitation each year.

    The Influence of Local Topography

    Rain-Shadow Effect

    Mountain ranges have a profound influence on climate. Figure 1.7 is an exaggerated cross section through the state of California at about the latitude of Fresno. It shows how the Sierra Nevada intercepts precipitation on its western slope. This phenomenon is known as the rain-shadow effect. As moisture-laden air flows inland from the ocean, it passes eastward, rising over the Sierra Nevada. Air is chilled adiabatically as it reaches higher elevations, like rising air at the equator. This chilling causes precipitation on the side of the mountain toward the coast. As the air descends from the mountain into Owens Valley, on the eastern side of the Sierra, it becomes heated by compression and causes evaporation. Thus, the Great Basin Desert occurs in the rain shadow of the Sierra Nevada.

    FIGURE 1.7 The influence of local topography on climate and weather, as indicated by an exaggerated cross section of California at about the latitude of Fresno.

    On the western side of the Sierra Nevada, the influence of elevation is similar to that of changes in latitude. As air moves up the western slope of the Sierra, its temperature decreases adiabatically by about 3–5°F/1000 ft (1°C/100 m), which is roughly equivalent to moving about 300 mi (480 km) northward. Precipitation increases accordingly, and the form of precipitation switches to snow. Moving northward, biotic zones occur at lower elevation. In northern California, coniferous forests occur at sea level. Climate above 11,000 ft (370 m) is roughly equivalent to that of the Arctic. A mountain is thus a microcosm of many types of climate.

    Temperature Inversion Layers

    Valleys also affect climate. Dense air flows downhill like water, a phenomenon known as cold-air drainage. Also like water, cold air collects in low spots that have limited or no drainage, such as the Great Central Valley of California. Cold air flows from the mountains at night and collects in the valley, filling it to the brim. In this case, the brim is the top of the Coast Ranges. Sometimes this cold air is so dense that winds are unable to disturb it. These circumstances produce dense fog that may remain in the valley for many days. In the Great Central Valley, this is known as a tule fog. Episodes of tule fog are famous for causing massive traffic jams and huge chain-reaction collisions. What is unique about these puddles of cold air is that normal temperature relationships seem to be inverted. It is supposed to become gradually cooler as one goes higher in an air mass, but when cold-air masses are trapped in a valley, it becomes suddenly warmer at the top. This paradoxical phenomenon is known

    Enjoying the preview?
    Page 1 of 1