Historic, Archive Document Do not assume content reflects current scientific knowledge, policies, or practices. ( s United States }j Department of Agriculture Forest Service Rocky Mountain Forest and Range Experiment Station Fort Collins, CO 80526 General Technical Report RM-258 Desired Future Conditions for Pinon-Juniper Ecosystems August 8-12, 1994 Flagstaff, Arizona Shaw> Douglas W; Aldon, Earl E; LoSapio, Carol, tech. coords. 1995. Desired Future Conditions for Pinon-Juniper Ecosystems; proceedings of the symposium; 1994 August 8-12; Flagstaff, Arizona. Gen. Tech. Rep. RM-258. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 226 p. Abstract The purpose of this symposium was to assist the USDA Forest Service, other federal land management agencies, and the Arizona State Land Office in managing pinon-juniper ecosystems in the Southwest. Authors assessed the current state of knowledge about the pinon-juniper resource and helped develop desired future conditions. Note: As part of the planning for this symposium, we decided to process and deliver these proceedings to the potential user as quickly as possible. Thus, the manuscripts did not receive conventional Forest Service editorial processing, and consequently, you may find some typographical errors. We feel quick publication of the proceedings is an essential part of the symposium concept and far outweighs these relatively minor distractions. The views expressed in each paper are those of the author and not necessarily those of the sponsoring organizations or the USDA Forest Service. Trade names are used for the information and convenience of the reader, and do not imply endorsement or preferential treatment by the sponsoring organizations or the USDA Forest Service. USDA Forest Service General Technical Report RM-258 Desired Future Conditions for Pinon-Juniper Ecosystems August 8-12, 1994 Flagstaff, Arizona Technical Coordinators: Douglas W. Shaw USDA Forest Service Southwestern Region Earl F. Aldon USDA Forest Service (Retired) Rocky Mountain Station Caroi LoSapio USDA Forest Service Rocky Mountain Station Rocky Mountain Forest and Range Experiment Station U.S. Department of Agriculture Fort Collins, Colorado Contents Welcoming and Opening Remarks Jose Salinas and Chip Cartwright 1 Western Juniper Woodlands: 100 Years of Plant Succession Rick Miller, Jeffrey Rose, Tony Svejcar, Jon Bates, and Kara Paintner 5 Soil Loss in Pinon-Juniper Ecosystems and Its Influence on Site Productivity and Desired Future Condition Malchus B. Baker Jr., Leonard F. DeBano, and Peter F. Ffolliott 9 Plant Species Composition Patterns with Differences in Tree Dominance on a Southwestern Utah Pinon-Juniper Site Robin J. Tausch and Neil E. West 16 Acid and Alkaline Phosphatase Dynamics in Soils of a Pinon -Juniper Woodland Susanne Kramer 24 Environmental Stress Influences Aboveground Pest Attack and Mycorrhizal Mutualism in Pinon-Juniper Woodlands: Implications For Management in the Event of Global Warming Catherine Gehring and Thomas Whitham 30 Stand Dynamics on Upper Elevation Pinon-Juniper Watersheds at Beaver Creek, Arizona Gerald J. Gottfried and Peter F. Ffolliott 38 Preliminary Results of Decomposition and Cellulose Degradation Along an Environmental Gradient in Northern Arizona Carole Coe Klopatek, Kenneth L. Murphy, Julie Rosen, John R. Obst, and Jeffrey M. Klopatek 46 Deer, Small Mammal, and Songbird Use of Thinned Pinon-Juniper Plots: Preliminary Results Steven K. Albert, Nelson Luna, and Albert L. Chopito 54 Relationships Among Plant Species Composition and Mule Deer Winter Range Use on Eastern Nevada Pinon-Juniper Chainings Robin J Tausch and Paul T. Tueller 65 Characteristics of Pinon-Juniper Habitats Selected for Feeding by Wintering Merriam's Turkey Brian F. Wake ling and Timothy D. Rogers 74 Wildlife Associations in Rocky Mountain Juniper in the Northern Great Plains, South Dakota Mark A. Rumble and John E. Gobeille 80 Effects of Fuelwood Harvesting on Small Mammal Populations in a Pinon-Juniper Woodland William H. Kruse 91 Cone and Seed Insects Associated with Pinon Pine Jose F. Negron 97 Insect and Disease Associates of the Pinon-Juniper Woodlands Terrence J. Rogers 107 Hydrology and Ecology of Pinon-Juniper Woodlands: Conceptual Framework and Field Studies Bradford R Wilcox and David D. Breshears 109 Understory Production and Composition in Pinon-Juniper Woodlands in New Mexico Rex D. Pieper 120 A Checklist for Ecosystem Management in Southwestern Pinon-Juniper Earl F. Aldon, Reggie Fletcher, and Doug Shaw 125 Woodland Inventory Procedures and Analyses Conducted for Management Planning Purposes on Indian Lands John Waconda 130 Watershed Restoration Through Integrated Resource Management on Public and Private Rangelands Sid Goodloe 136 "Can't We All Just Get Along" Jon S. Bumstead 141 Responding to Tribal Voices in Managing Woodland Resources Ronald K. Miller 146 The Capulin Pinon-Juniper Ecosystem Management Project The Archaeological and Ecological Components John C. Phillips and Martha D. Yates Ph.D 153 Community Based Pinon-Juniper Woodland Resource Management Planning for the Nahaf a' Dziil Chapter Usha Little and Denver Hospodarsky 160 An Overview Of Woodland Projects Incorporated At Four Pueblos In New Mexico Buff Jebsen-Ross and Richard Schwab 168 The Effects of Fire on Cultural Resources Mesa Verde National Park, Colorado Kathleen Fiero 176 Western Juniper: An Evolving Case Study in Commercialization, Ecosystem Management, and Community Development Larry Swan 179 Tres Piedras Pinon-Juniper Silviculture: A Partnership Project Between the USDA Forest Service and New Mexico State University John T. Harrington, Jim Fitch, and Patrick A. Glass 184 Pinon Pine Seed Production, Collection, and Storage Richard M. Jeffers 191 Carrizo Demonstration Area Restoration of a Southwest Forest Ecosystem Richard S. Edwards 198 Silvicultural Systems for Pinon-Juniper James R. Ellenwood 203 Trial Applications of Low-Impact Herbicides for Pinon-Juniper Control in the Southwest Douglas Parker, Max Williamson, Richard Edwards, and Russell Ward 209 Pinon-Juniper Fuelwood Markets in the Southwest Lawrence A. Schmidt 214 Ecosystem Management Research in an "Old Growth" Pinon-Juniper Woodland William H. Kruse and Hazel M. Perry 219 The Composition of Oils in Pinus edulis Michael Blair, Telletha Valenski, Andrew Sykes, Russell Balda, and Gerald Caple 225 Welcoming and Opening Remarks Jose Salinas1 and Chip Cartwright2 Jose Salinas... Good Morning! I am pleased to join the rest of the speakers this morning in welcoming each of you to this week's 1994 Pirion-Juniper Symposium — Desired Future Conditions for Pinon-Juniper Ecosystems. I, too, want to thank our co-sponsors: • Northern Arizona University • University of Arizona • Bureau of Indian Affairs, Phoenix Area Office • Society of American Foresters Equally important, I want to thank Doug Shaw and the Steering Committee that coordinated this symposium Doug has been coordinating the P-J Initiative for a little over two years. I want to thank the other agency coordinators also. There are always behind-the-scene players that are part of a successful program I want to recognize my Regional office watershed staff: • Penny Luehring • Debby Potter • Chic Spann • Wayne Robbie • Penny Fabian • Barney Lyons And two staff people who helped with symposium but stayed in Albuquerque to keep the office in operation — Gena Velasquez and Theresa Sanchez. A key partner and supporter in the P-J watershed restoration program is Gerald Henke, Regional Director for Range Management. Of course, we are very pleased to have our Regional Forester, Chip Cartwright, with us this morning. Chip will be properly introduced in a moment I also want to thank the Coconino National Forest for hosting us as well as for hosting our field trip. I understand more than 200 people signed up for this symposium. I think this in itself is a success and we have not even begun! 1 Director, Watershed Management and Air Management, USDA Forest Service, Southwestern Region, Albuquerque, NM. 2 Regional Forester, USDA Forest Service, Southwestern Region, Albuquerque, NM. We have people representing several Regions. Our presenters include research, Forest Service managers, program leaders, people with specific expertise, academia, members of other agencies, people representing tribal governments, members of the ranching community, and private industry. And the list goes on and on. The agenda consists of scientific papers, papers on human and social relationship within the P-J ecosystem We also have a series of presentations of on going projects. Plus, we have several posters on display and a couple of videos on P-J. Based on the information I have shared with you, I think all of you will agree with me that the title of the symposium — Desired Future Conditions for Pinon-Juniper Ecosystems — is very appropriate. So why are we all here this week? For one thing, not all is well with our P-J woodland ecosystems. There are over 7 million acres of P-J on national lands in the Southwestern Region alone. But the real concern is that nearly 50% of this land base represents watersheds in unsatisfactory condition. This condition is not unique to this region nor to the Forest Service. For the same condition exists on other federal lands, Indian land and private lands. This condition also extend to all the states surrounding us, including Mexico. Our P-J ecosystem was never in this type of condition — in other words, these unsatisfactory conditions are not natural. We, man, through our past management practices, have created unbalance among the components of the ecosystem — the biological, the physical and the human aspects. We can no longer continue to contribute to the problem. We must work together — the research scientist, the land manager and the user — be state, federal or private, we must work together in seeking solutions. This may include joint projects under partnership or it could mean technology transfer. It may include changing present land management practices. It will mean being more responsive to the needs of the land and the people who are part of that ecosystem. 1 Definitely, a scientific-based ecosystem manage- ment program is the appropriate approach. Deteriorating watersheds are not a problem that belongs to the project leader in a remote Ranger District. It is your problem. It is my problem. It is a social and economic problem. It is our problem — yours and mine together. I am proud to announce that the Southwestern Region has increased its funding for watershed restoration work in the last two years by more than 38%. I am proud to announce that every Forest in the Region has a P-J demonstration project. And I am proud to announce nearly all of the 69 Ranger Districts in the Region have identified some kind of watershed restoration demonstration project. There has been a tremendous amount of commit- ment by our field personnel and our partners. Our management practices are being reassessed to be in balance with the watershed capability. We are making better use of our soil inventory data as well as research data. We are all doing a good job — but we must do better if we are to get in front of the many pressures, problems and issues facing our P-J ecosystems. When it is all done, water quality and soil productivity will be the determining factors to the health and well being of our nation or any nation. We have a lot at stake if we are to remain a strong nation and one of leadership and balance between people and the environment. If I could have only one outcome from this symposium, it would be that our partnerships in P- J watershed restoration would increase in number and strength. These partnerships could be based on technology transfer, research, joint projects, and promotion of ecosystem management principles. In summary, I encourage each of you to participate in symposium as much as possible. And to have an enjoyable and productive week. Thank you. ****** Chip Cartwright... Welcome to this 1994 symposium on Pinon- Juniper management. I'm glad you are here to share information about these important southwestern ecosystems. Being new to the Southwestern Region, I'm glad to be here to learn a little about how we are implementing ecosystem management principles in the Pinon-Juniper. I want to stress a few points about ecosystem management in these brief opening comments. First ecosystem management as we know it is adaptive to changes in our knowledge about the system. Second humans and their interactions with the other ecosystem components are part of ecosystem management, and we are stressing involvement of all stakeholders in management. Third, ecosystem management is based on sound science, and forth ecosystem management must involve strong partnerships. The Pinon-Juniper is vast; it stretches in some form from west Texas to the Santa Ynez Mountains of Southern California and from Oregon deep into Mexico. The story of these woodlands extends back through historic, prehistoric, and geologic time. The history of the Pinon-Juniper is a documen- tation of humanities' changing relationship with its environment. It starts with a relationship in which human fate was determined by the bounty of its un-manipulated environment, and extends through relationships in which people dramatically modify their environment to meet the needs and wants of growing populations. The purpose of this symposium is to share information that will be useful in directing human relationships with this ecosystem in the future. Our goal for this relationship should be ecosystem sustainability. We understand no ecosystem is a simple thing, but P-J, with so many links to people and cultures, is particularly complicated. Ron Lanner describes this complexity well in the introduction to his book on the Pinon Pine, "Every tree, like every other living organism, is at the center of its own four-dimensional spider web: tug on this strand or that and see what quivers, what falls, what comes in or goes out, what lives or dies, what grows fat-and when." Much has been said in the last decade or so about managing Pinon-Juniper. Bill Hurst, Regional Forester from 1966-1976, said there were more symposiums about-but less done about — P-J, than on any other species he could remember. Why do we continue to talk a lot and act little? Perhaps it is because we don't understand or take seriously the importance of this ecosystem to current human needs and past and current human beliefs and perceptions. Or perhaps we cannot reach a common agreement on how to manage an ecosystem that is so highly variable and has so many ties to so many different cultures. Societal values and perspectives about the environment change and evolve continuously. The human domination era, of most western societies, was replaced early in the 20th century by the conservation and multiple-use era. Some societal values and perceptions are shifting again, as we 2 enter the 21st century. In response to this shift the Forest Service is looking more holistically at natural systems as systems that integrate biological, physical, and human dimensions. We have a heightened concern for the long term sustainability of systems, not just the multiple-uses that can be realized from these systems. This evolving approach to land stewardship is frequently referred to as "ecosystem management". Along with the evolution of attitudes about ecosystem management is an increased sensitivity to the importance of stakeholders in the management of public lands. The Region recently completed work to help us define principles for ecosystem management. A human dimension study group is publishing a short document containing eight human dimension principles for ecosystem management. A scientific study team recently published a document titled, "An Ecological Basis for Ecosystem Management" which contains six guiding principles for ecosystem management. Both documents have the following common principle, "Humans are an integral part of today's ecosystems and depend on natural ecosystems for survival and welfare; ecosystems must be sustained for the long-term well-being of humans and other forms of life". In the past, our efforts to assess the physical/biological aspects of ecosystems have been separate from our efforts to assess the human dimension aspects of ecosystems. We are working hard to understand how these two aspects can be integrated. According to Eugene Odum, "Ecology is the study of the structure and function of nature, it being understood that mankind is part of nature". The culture, family heritage, lifestyles and livelyhoods of people are linked with the ecosystems of which they are a part. Peoples past, present, and future values and desires influence ecosystems. Ecosystems affect people's physical, mental, spiritual, social, cultural, and economic well-being. Understanding these relationships establishes a basis from which integrated ecosystem management can contribute to sustaining human life as well as healthy ecosystems. As I mentioned earlier we are becoming much more sensitive to the needs of stakeholders in our efforts to implement ecosystem management. How do we integrate the needs of such a diverse spectrum of people into an ecological approach to Pinon-juniper management. Some people consider the P-J "public enemy number one", while others consider the P-J as "our very life! They are part of us." Still others have a long tradition of using P-J for daily needs such as cooking and heating. We also have the needs of the biological part of this system to understand, how much can we harvest with out adversely affecting the needs of the Pirion jay? And the physical part must be integrated. How do we protect or restore soil productivity, a basic consideration to all other needs? We must look to the people for answers to these and many other questions. We know people want to be involved, but we also know people must be sensitive and knowledgeable to be effectively involved. It is a basic human dimension principle that sustainable ecosystem management requires an ecologically knowledgeable population. We also look to scientists for answers. Scientific information helps us better understand the range of choices for actions and the consequences of following one path instead of another. Scientists can provide the citizen-manager partnership with some of the information required to make informed decisions. Research will assist in better understanding ecosystem functions and inter- relationships and all scales and times. Research will also assist in better understanding stakeholders and the values and motives that drive their behavior. We must seek a more complete understanding of the social and cultural dimensions of ecosystems, including the changing perceptions, needs, and values of people. Adaptive management is a term currently used to describe an evolving management strategy for ecosystems. In such a strategy ecosystems like the P-J and related ecosystems will be managed according to a working hypothesis which continues to change and adapt to new information and experience. To make this management strategy effective we must monitor and evaluate management to see if we are achieving the desired results and if the results we originally desired are still desired. The P-J is vast and the problems are wide spread in the Southwest. We have goals for the National Forest part of P-J that include maintaining or improving soil productivity, meeting water quality standards, maintaining or improving visual and biological diversity, protecting habitat needs for threatened and endangered species, and historic and prehistoric cultural values. But our goal also includes management that is sensitive to lifestyles as well as ecosystem needs. But the P-J involves more than just National Forest. All the stakeholders in this vast system must join in partnerships to describe and move toward shares 3 visions and goals. We are actively seeking such partnerships. We have a partnership with two great universities, a professional society, and another agency to organize this symposium. But our partnerships must expand to include others that have entirely different visions for the P-J. Our partnerships must change to collaborative partnerships with shared visions and also shared responsibilities for bring those visions into realization. I hope this symposium is successful in sharing the latest in scientific information. I also hope this symposium is successful in bring together stakeholders in a way that will lead to partnerships and most important sustainable ecosystem management in the P-J. 4 Western Juniper Woodlands: 100 Years of Plant Succession Rick Miller1, Jeffrey Rose2, Tony Svejcar3, Jon Bates4, and Kara Paintner2 Abstract. — During the past 100 years there have been dramatic increases in western juniper (Juniperus occidentalis spp. occidentalis) in the western U.S. Evidence for the increase come from descriptions of explorers and early settlers, old photographs, ring counts of existing trees, and pollen cores taken from pond sediments. A number of factors may have contributed to the western juniper expansion. Most probable among the contributing factors are: 1) reduction or removal of Native American populations that actively used fire as a vegetation management tool; 2) removal of fine fuels, and thus a decline in fire frequency as a result of heavy livestock grazing between 1880 and 1930; and 3) mild temperatures and above average precipitation during the late 1880s and early 1900s. Because young juniper (<50 yr. old) do not survive fires, most factors causing a reduction in fire frequency will tend to favor western juniper. INTRODUCTION One of the most pronounced changes in plant community dynamics in the western U.S. has occurred in the juniper and pihon-juniper woodlands, a major vegetation type characterizing the Intermountain Region. These woodlands, sometimes described as pygmy forests, currently occupy 17 million ha throughout this region (West 1988). Juniperus occidentalis ssp. occidentalis Hook (western juniper) is considered the Northwest representative of the pihon-juniper zone in the Intermountain Region (Franklin and Dyrness 1973) and occupies over 1 million ha (Dealy et al. 1978) in eastern Oregon, southwestern Idaho and northeastern California (Cronquist et al. 1972). This subspecies of /. occidentalis is found primarily north of the polar front gradient (Neilson 1987) (parallel to the Oregon and Nevada border, Latitude 42°) where temperatures are cooler, summer precipitation decreases, and winter precipitation increases (Mitchell 1976). Relict juniper woodlands, tree-age class ratios and historical documents generally indicate western juniper woodlands, prior to Euro- American settlement, were open, sparse and savannah-like (Burkhardt and Tisdale 1969, Vasek and Thorne 1977, Miller and Rose 1994, EOARC data files). In southeast Oregon and northeast California, where soils are primarily derived from igneous material, the majority of old trees established prior to settlement are located on the shallow soil low sagebrush flats and rocky ridges where fine fuels were insufficient to carry a fire (Vasek and Thorne 1977, EOARC data file). Densities of presettlement trees on these harsh sites generally ranged from 8 to 20 per ha (EOARC data file). Old growth stands in the ash-pumice zone of Mt. Mazama and Newberry Crater in central Oregon are more extensive (personal observation by authors). During the last 100 years, western juniper has increased both in distribution and density throughout its range. Expansion has been most dramatic in the deeper more productive 1 Range Scientist, Eastern Oregon Agricultural Research Center, Oregon State University, Burns, OR. 2Research Assistant, Eastern Oregon Agricultural Research Center, Oregon State University, Burns, OR. 3 Supervisory Range Scientist, USDA Agricultural Research Service, BurnSj OR. Graduate Research Assistant, Department of Rangeland Resources, Oregon State University, Corvallis, OR. 5 soils of open meadows grasslands, sagebrush steppe communities, aspen groves, and riparian communities (Eddleman 1987, Miller and Rose 1994, Miller and Wigand 1994, Young and Evans 1981). In central Oregon in 1825, Ogden observed only occasional junipers (reported as cedars) growing on the hillsides, while traveling through the Crooked River drainage in central Oregon (Rich et al. 1950). Today, these hillsides are covered by dense juniper woodlands. In a nearby area, J.W. Meldrum's 1870 survey notes describe a gently rolling landscape covered with an abundance of perennial bunchgrasses and a wide scattering of juniper trees (Caraher 1977). Today, juniper densities on this site vary between 125 and 250/ha. Near Silver Lake Oregon, juniper density increased from 17/ha in 1890 to over 400/ha by 1970 (Adams 1975). The majority of trees established between 1902 and 1936. On another site in central Oregon where junipers were absent prior to 1880, densities reached 1018/ha by 1980 (Eddleman 1987). David Griffiths, a representative from the Department of Agriculture, was sent to tour and evaluate the condition of the western rangelands around the turn of the century. He observed only scattered stands of juniper on Steens Mountain in southeastern Oregon (Griffiths 1902). Western juniper began increasing in both density and distribution in the late 1800s in central and southeastern Oregon, northeast California, and southwest Idaho (Eddleman 1987, Miller and Rose 1994, Young and Evans 1981). In much of central Oregon and portions of northeastern California, peak establishment occurred between 1890 and 1930. In southeastern Oregon, juniper expansion began in the 1880s, however, establishment progressed slowly during the early 1900s; the rate increased later into the century as trees reached reproductive maturity (Eddleman 1984, Miller and Rose 1994). Current densities of trees less than 100 years old on the more productive low sagebrush and mountain big sagebrush communities on Steens Mountain average 338/ha (Miller and Rose 1994). In stands where juniper has invaded and completely replaced aspen trees on Steens Mountain, densities of mature juniper trees ranged from 725 to over 2,000/ha (Miller and Rose 1994). The oldest juniper trees in these stands were less than 90 years old. Although western juniper is a long-lived species, the oldest tree reported in Oregon is 886 years old (Holmes et al. 1986). The majority of present day woodlands in eastern Oregon are less than 100 years old (USDI-BLM 1990). Seed dissemination occurs primarily through movement by water across the land surface, particularly on frozen soils (Eddleman 1984), and through bird, coyote and rabbit dispersal (Gabrielson and Jewett 1970). The Townsend solitaire (Myadestes townsendii) (Lederer 1977, Poddar and Lederer 1982), the American robin (Turdus migratorius), Steller's jay (Cyanocitta stelleri) and scrub jay (Aphelocoma coerulescens) are primary avian vectors of juniper seed dispersal in the Great Basin (Gabrielson and Jewett 1970). The factors most frequently associated with the recent expansion of juniper species throughout the west are climate, fire, and grazing. The combined affect of climate and fire were likely the cause of juniper expansion and retraction during prehistoric times (Miller and Wigand 1994), but are climate change and altered fire regimes fully responsible for the expansion of western juniper woodlands during the last 100 years? Following the end of the Little Ice Age in the mid 1800s, winters became more mild and precipitation increased above the current long term average between 1850 to 1916 in the northern Great Basin (Antevs 1938, Graumlich 1985, Holmes et al. 1986). Mild conditions and increased precipitation during the late 1800s and early 1900s, which promotes vigorous juniper growth (Earle and Fritts 1986, Fritts and Xiangdig 1986), increased the potential for juniper establishment. In central Oregon, juniper establishment primarily occurs during years of good tree ring growth, with few trees establishing in years of marginal tree ring growth (Adams 1975). However, these conditions would also have increased the potential for fire due to the increased production of fine fuels; grasses and forbs. Reduced fire frequency has been one of the factors attributed to the expansion of juniper throughout the west (Burkhardt and Tisdale 1976, Young and Evans 1981). Before settlement, fire frequencies in mountain big sagebrush (Artemisia tridentata spp. vaseyana) communities varied from 15 to 25 years (Burkhardt and Tisdale 1976, Martin and Johnson 1979, Houston 1982). Western juniper less than 40 to 50 years old are easily killed by fire (Burkhardt and Tisdale 1976). Fire probably maintained both shrubs and trees at low densities and often restricted trees to the harsh sites which produced little contiguous fuel. Reduction of fire frequencies during settlement was probably due to a decline in Native American- set fires and the reduction of fine fuels through livestock grazing. The effects of fire suppression during the late 1800s and early part of this century were minimal, not becoming a factor until after WWII. Native American-caused fires augmented Hghtning fires in the more mesic sagebrush communities (Agee 1993). Fire was used to improve forage for game, maintain or increase the yield of certain wild edible plants and increase seed production. However, the influence of Native Americans declined as early as the late 1700s. By the close of the eighteenth century, native populations throughout the Intermountain Region were reduced 80 percent by European diseases such as smallpox, measles, venereal disease and possibly typhus (Thompson 1916, Cressman 1981). Despite their decline in population, Peter Skene Ogden noted abundant evidence of Native American set-fires in the Harney and Malheur lakes region during the middle 1820s. Settlement of the region by European Americans in the late 1800s and early 1900s, lead to a reduction of fine fuels through grazing high densities of domestic livestock (Griffiths 1902, Burkhardt and Tisdale 1976). Possibly the greatest influence livestock had on the expansion of juniper throughout the West was the reduction of fine fuels resulting in a decrease in fire return intervals. In 1901, on his trip from Nevada to eastern Oregon, Griffiths (1902) stated; "No open-range lowland was seen on the whole trip which had much feed upon it excepting that consisting of the tough and persistent salt grass. On the whole trip of three days we found no good feed, except in very steep ravines/' Removal of fine fuels was probably of particular importance during the wet and mild climate conditions of the late 1800s and early 1900s initiating the development of western juniper woodlands. Competition appears not to be a factor inhibiting western juniper seedling establishment (Burkhardt and Tisdale 1976, Eddleman 1987, Miller and Rose 1994). Ecological condition of a plant community does not appear to influence seedling establishment. However, an increase in sagebrush would increase the number of safe sites for juniper seedling establishment. The majority of juniper seedlings are usually found beneath sagebrush canopies (Burkhardt and Tisdale 1976, Eddleman 1984, Miller and Rose 1994). A more recent argument attributes the expansion of pinon-juniper woodlands in the southwest to increased atmospheric CO2 concentrations (Johnson et al. 1990). Bazzaz et al. (1985) reported cool season (C3 ) plants respond more favorably to increased C02 levels than do warm season (C4 ) plants. In the southwest, increased atmospheric CO2 may increase growth of C3 junipers at the expense of associated C4 grasses in the understory. In the northern portions of the juniper zone (e.g., western juniper) understory species are also C3 forbs and grasses. However, water use efficiency has been shown to be enhanced more in woody than in herbaceous cool season plants (Polley et al. 1993). CONCLUSION Since the turn of the century western juniper has rapidly increased in distribution and density. Although western juniper has fluctuated during the prehistoric past, several factors appear to be different between past and present expansion (Miller and Wigand 1994). The present expansion differs from past increases, in that it has occurred during a period of increasing aridity. In the past, increasing aridity generally increased the fire frequencies and caused a decline in western juniper dominance. During the past 70 years aridity has increased, but human activities have greatly reduced fire frequencies. LITERATURE CITED Adams, A.W. 1975. A brief history of juniper shrub populations in southern Oregon. Oregon State Wildlife Commission. Wildlife Res. Rep No. 6. Agee, J.K. 1993. Fire ecology of Pacific Northwest forests. Island Press, Washington, D.C. Antevs, E. 1938. Rainfall and tree growth in the Great Basin. Carnegie Instn. of Wash., Publ. 469, Am. Geogr. Soc, Spec. Publ. 21. Bazzaz, FA., K. Garbutt, and WE. Williams. 1985. Effects of increased atmospheric carbon dioxide concentration on plant communities. Pages 155-170. In: B.R. Strain and J.D. Cure, eds. Direct effects of increasing carbon dioxide on vegetation. US Dept. of Energy, DOE/ER-0238. Burkhardt, J.W, and E.W Tisdale. 1976. Causes of juniper invasion in southwestern Idaho. Ecology 76:472-484. Caraher, D.L. 1977. The spread of western juniper in central Oregon, p. 3-8. In: R.E. Martin, J.E. Dealy and D.L. Caraher (eds.), Proc. Western juniper ecology and management workshop. USDA For. Serv. Gen. Tech. Rep. PNW-4. Cressman, L.S. 1981. The Sandal and the Cave. Oregon State University Press, Corvallis, Oregon. Dealy, J.E., J.M. Geist, and R.S. Driscoll. 1978. Western juniper communities on rangeland of the Pacific Northwest. Pages 201-204. In: D.E. Hyder ed., Proceedings to the First International Rangeland Congress, Denver, CO. 7 Earle, C.J., and H.C. Fritts. 1986. Reconstructing river flow in the Sacramento Basin since 1560. Rep. Calif. Dept. Resources, Agreement No. DWR B-55395. Lab. of Tree-ring Res., Univ. Ariz., Tucson. Eddleman, L.E. 1984. Ecological studies on western juniper in central Oregon, p.27-35. In: T.E. Bedell (ed.), proc. western juniper management short course, Oregon State Univ. Ext. Serv. Corvallis, OR. Eddleman, L.E. 1987. Establishment and stand development of western juniper in central Oregon. Pages 255-259. In: R.L. Everett , ed. Proc. Pinyon- Juniper Conf. US Forest Service General Technical Report INT-215. Franklin, J.F., and C.T. Dyrness. 1973. Natural vegetation of Oregon and Washington. USDA Forest Service, General Technical Report PNW-8. Portland, OR. Fritts, H.C., and W. Xiangdig. 1986. A comparison between response-function analysis and other regression techniques. Tree-ring Bull. 46:31-46. Gabrielson, I.N., and S.G. Jewett. 1970. Birds of the Pacific Northwest. Dover Publ. Graumlich, L. 1985. Long-term records of temperature and precipitation in the Pacific Northwest derived from tree rings. PhD. Thesis, University of Washington, Seattle. Griffiths, D. 1902. Forage conditions on the northern border of the Great Basin. Bureau of Plant Industry. USDA, Bull 15. Holmes, R.L., R.K. Adams, H.C. Fritts. 1986. Tree-ring chronologies of western North America: California, eastern Oregon and Northern Great Basin. Laboratory of Tree-Ring Research, University of Arizona, Chronology Series VI. Houston, D. B. 1973. Wildfires in northern Yellowstone National Park. Ecology 54:1109-1117. Johnson, H.B., S.S. Mayeux, Jr., and H.W Polley. 1990. Increasing atmospheric C02 concentrations and vegetation change on rangelands. Proc. Soc. Range Manage. 43rd Annual Meeting. Reno, Nevada. Lanner, R.M. 1984. Trees of the Great Basin: A Natural History. University of Nevada Press. Reno. Lederer, R.J. 1977. Winter territoriality and foraging behavior of the Townsend's Solitaire. Amer. Midi. Natur. 97:101-109. Martin, R.E. and A. Johnson. 1979. Fire management of Lava Beds National Monument. In: Linn, R.M. ed., Proc. of the first conference on scientific research in the nation parks: pp. 1209-1217. USDI Nat. Park Serv. Trans, and Proc. Ser. 5. Miller, R.F. and PE. Wigand. 1994. Holocene changes in semiarid pinyon-juniper woodlands: response to climate, fire, and human activities in the US Great Basin. Bioscience 44:465-474. Miller, R.F. and J.A. Rose. In press. Western juniper expansion in eastern Oregon. Great Basin Nat. Poddar, S., and R.J. Lederer. 1982. Juniper berries as an exclusive winter forage for Townsend's Solitaires. Am. Midi. Nat. 108:34-40. Polley, H.W, H.R. Johnson, B.D. Marino, and H.S. Mayeux. 1993. Increase in C3 plant water-use efficiency and biomass over glacial to present C02 concentrations. Nature 361:61-64. Rich, E.E., A.M. Johnson, and B.R. Baker (eds.). 1950. Peter Skene Ogden's Snake Country Journals: 1824- 25 and 1825-26. The Hudson Bay Society, London. Thompson, D. 1916. David Thompson's narrative. J.B. Tyrrel, (ed.), The Champlain Society, Toronto, Ontario, Can. Vasek, EC. and R.F. Thorne. 1977. Transmontane coniferous vegetation. Pages 797-832. In: M.G. Barbour and J. Major, eds. Terrestrial Vegetation of California. California Native Plant Society. Spec. Publ. No. 9. West, N.E. 1988. Intermountain deserts, shrub steppes, and woodlands. Pages 209-230. In: M.B. Barbour and WD. Billings eds., North American Terrestrial Vegetation. Cambridge University Press. Young, J.A., and R.A. Evans. 1981. Demography and fire history of a western juniper stand. J. Range Manage. 34:501-506. 8 Soil Loss in Pinon- Juniper Ecosystems and Its Influence on Site Productivity and Desired Future Condition Malchus B. Baker Jr.1, Leonard F. DeBano2, and Peter F. Ffolliott3 Abstract. — Pinon-juniper woodlands are widespread throughout the western United States and have provided habitat and a variety of products for human use in recent and historical times. Site productivity of pinon- juniper woodlands depends on a wide range of complex processes which dynamically interact over various time and space scales. The hydrology of pinon-juniper ecosystems is relatively complex, because it is controlled by interactions among precipitation regimes, geomorphological settings, and edaphic conditions. Superimposed on the natural system is a wide spectrum of past land uses, and misuses. The present and past hydrologic environment has also been characterized by extreme spatial and temporal variation. Water and wind erosion are primary processes that influence site productivity. In this paper, a limited amount of experimental data on water and wind erosion is presented for southwestern pinon-juniper woodlands. A conceptual model is presented for describing water and wind erosion and is used to illustrate their dependency on mean annual precipitation for vegetated and disturbed (bared) sites. This model emphasizes the importance of plant cover density on soil loss and subsequent site productivity. Soil loss is a crucial factor affecting productivity because nutrient enriched materials are lost from the site during water and wind erosion. Decreases in site productivity resulting from soil loss make it more difficult to attain ecological goals necessary for achieving different desired future conditions. It is imperative, therefore, that management practices implemented on pinon-juniper woodlands minimize soil losses and associated reductions in site productivity. INTRODUCTION pinon-juniper woodlands, similar to those found throughout the Southwest, are one of the most extensive vegetative type in the western United States (Evans 1988). These woodlands provide a wide range of valuable products and benefits such as fuelwood, fence posts, Christmas trees, pinon nuts, forage for livestock grazing, and critical habitat for a large number of game and nongame animals, including some rare and ' Research Hydrologist, USDA Forest Service, Forestry Sciences Complex, Flagstaff, AZ. 2 Research Soil Scientist, USDA Forest Service, Tucson, AZ. Professor Watershed Management, University of Arizona, Tucson, AZ. endangered species. Some products produced in these woodlands (e.g. pinon nuts and firewood) have played a significant role in sustaining past human occupancy and cultures, as well as providing useful products to present day society. Site productivity of pinon-juniper woodlands depends on a wide range of complex processes which interact dynamically over time and space. Although site productivity can be viewed simply as the capacity of a soil to support plant growth (Powers 1991), this concept encompasses several 9 complicating factors, including the effects of soil erosion. Because soil erosion historically was a common occurrence in pihon-juniper woodlands (both naturally and in response to human activities), it is important to understand how these losses may affect site productivity and, ultimately, the overall sustainability of these woodlands. Equally important are the effects that changes in site productivity may have on the desired future condition (both physio -biological and socio-economic dimensions) of these woodlands. The objectives of this paper are to: (1) present a conceptual framework for describing erosion processes; (2) discuss the impacts of soil losses on site productivity; and (3) to relate erosion and site productivity information to expected changes in the desired future condition of southwestern pifion-juniper woodlands. CHARACTERISTICS OF PINON-JUNIPER WOODLANDS Physical and Biological Environment Pinon-juniper woodlands in the Southwest are found on foothills, low mountains, mesas, and plateaus between 1,300 and 2,200 meters elevation (Brown 1982, Clary et al. 1974, Gottfried 1992). These woodland communities occupy elevations between the more xeric lower elevation brush and grass-dominated vegetation types and more mesic higher elevation montane forests. Pifion pine tends to become dominate at higher elevations, whereas, junipers are more common at the lower elevations. The distributions of pinon-juniper woodlands reflect ecological amplitude and responsiveness to available soil water and temperature regimes, rather than to any general topographic relationship. The stand structure of pinon-juniper woodlands is relatively simple. In general, undisturbed stands are uneven-aged. Pifion tends to dominate the smaller size classes in these stands, while junipers are the major component of the larger size classes. Even-aged stands frequently develop as a result of fire and tree removal operations. The pifion-juniper ecosystems, however, have very heterogeneous overstory-understory relationships. Moir and Carleton (1987) recognized over 70 habitat types in Arizona and New Mexico pifion-juniper woodlands. Pifion-juniper woodlands are found on a variety of soils that are derived from diverse parent materials including intrusive and extrusive igneous rocks, sedimentary rocks, and mixed alluvium (Springfield 1976). Soil depths vary from shallow to moderately deep, and soil textures range from coarse, rocky, and porous sandy loams to fine compacted clays. Fertility levels of the soils are generally low to moderate. Hydrologic Environment The hydrology of pifion-juniper ecosystems is complex, because it reflects a wide range of interactions among precipitation regimes, geomorphological settings, and edaphic conditions. It is further complicated by a wide spectrum of past land uses, and misuse. Hydrologic scenarios vary from a desirable combination of low-intensity rainstorms, good vegetative cover, and permeable soils, to a potentially hazardous situation on areas with steep slopes, sparse vegetative cover, and impermeable soils that are often subjected to high-intensity rainstorms. A heterogeneous vegetative cover makes it difficult to generalize watershed characteristics and potential hydrologic functioning; and variations in soils, with their different infiltration rates, further complicate the hydrology. Climatic variability has also been a key factor affecting past and current community dynamics in pinon-juniper woodlands (Betancourt et al. 1993). Paleoecological data collected on a wide range of sites throughout the Southwest suggest that droughts have been a common and regular occurrence during the last 40,000 years. Measured precipitation in pinon-juniper woodlands range from 300 to 440 mm, with local areas receiving 500 mm, or more (Hibbert, 1979). Summer convective storms can account for over half of the total annual precipitation. Erosion and Soil Loss Soil losses by both water and wind erosion are an integral part of pinon-juniper ecosystems because precipitation required for a dense protective vegetative cover is marginal. This delicate balance between erosional processes and the conditions required for an acceptable vegetative cover make these ecosystems particularly sensitive to both natural disturbances 10 and improper land use practices. Imbalance can reduce vegetative cover and accelerate soil loss. Because a well maintained plant cover reduces soil losses, it is a cornerstone of land management activities. Most storms cause little or no overland flow from sites having good ground cover, gentle slopes, and permeable soils (Baker 1986, Clary et al. 1974, Springfield 1976). However, high- intensity, short-duration storms can produce significant runoff events, particularly when they occur on steep slopes having a sparse ground cover and soils with low permeability. Historically, high rates of runoff and sediment have been attributed to overgrazing by livestock, fire, or other past misuse (Evans 1988; Wood and Javed 1992). Water erosion on pifion-juniper woodlands in the arid Southwest generally can be characterized as an unsteady, or episodic process, that transports sediment from a source area, across a landscape, and through a channel system with intermittent periods of storage (DeBano 1977). Results obtained from studies in the Southwest provide some information on the relative magnitudes of soil losses associated with different management activities. For any site condition, runoff, erosion, and sediment production are related to precipitation events that move intermittently stored material downstream. After 9 years of pinon-juniper watershed studies on basalt-derived soils in central Arizona, the largest sediment yield recorded was 2200 kg/ha from a six-year-old, cabled watershed during an intense rainstorm — estimated recurrence interval of 100 to 150 years (Clary et al. 1974). Based on the knowledge of the influence of treatment intensities, age since treatment, and storm frequencies on sediment losses, it was concluded that maximum potential sediment yields were in the range of 2240 to 4880 kg/ha/year for watersheds with similar physical characteristics and climatic regimes (Clary et al. 1974). Runoff studies on small plots, using both natural and simulated rainfall, have provided additional information on the amounts of sediment moved during runoff from sites receiving different treatments. Under natural rainfall events, course sediment losses from plots where slash was piled and burned exceeded those of control plots by about four-fold (100 kg/ha from controls compared to over 400 kg/ha on burned slash plots) (Wood and Javed 1992). Loping and scattering slash without burning it had little effect on sediment production. Similarly, sediment concentrations were lower in runoff water from control plots than from those that had been burned. Total sediment yields from dry soil rainfall simulator runs (rainfall simulation on dry soil) on pihon-juniper sites in New Mexico and Arizona ranged from 368 kg/ha (per simulation) on high cover plots to 2211 kg'ha on plots that had been scraped bare (Bolton et al. 1992). Significant movement of soil by wind can also occur in pihon-juniper woodlands, particularly following prescribed burning of fuelwood slash. Mean sediment amounts collected in samplers located 1 meter above the ground during a summer season (May 15 through October) were as high as 17.9 kg/m2 (Baker and Jemison 1992). At 0.05 m above the soil surface as much as 1164 kg/m of soil and ash material were collected. These measurements of windborne material represent quantities of sediment caught at points in a vertical profile on the site sampled and not the total amounts that were removed from the site. A CONCEPTUAL FRAMEWORK FOR SOIL LOSS A model initially developed by Marshall (1973) and later extended by Heathcote (1983) to describe water and wind erosion can be used to illustrate erosional processes in pinon-juniper woodlands (Figure 1). This model emphasizes the dependency of water and wind erosion on mean annual precipitation for vegetated and bare sites. It also provides a useful framework for discussing the nature of wind and water erosional processes and their controlling factors. In figure 1, curve labeled waterveg represents water erosion under natural vegetation cover — normal erosion (Marshall 1973). Water erosion increases from a low value at the arid extreme of mean annual precipitation to a peak erosion rate in the semi-arid rainfall range (400 mm). Here, rainfall is not great enough to sustain a complete vegetation cover all year, but is sufficient to cause erosion of the bare soil. With additional precipitation, vegetation cover increases and water erosion decreases (Schumm and Harvey 1982). The curve labeled waterbar represents water erosion rates in the absence of natural vegetation, e.g. vegetation loss due to overgrazing or burning. Here, erosion continue to increase in excess of normal erosion rates (curve waterveg) in the humid range. The opportunity for reducing water erosion below the maximum rate (curve 11 waterbar) is quite low at the arid and semi-arid end of the scale, but increases markedly at the humid end. In contrast, normal wind erosion, curve windveg, decreases exponentially as precipitation increases because the presences of even a moderately sparse vegetation cover can reduce the wind force at the soil surface. Wind erosion, in the absence of vegetation (curve windbar)/ remains at a relatively high level until enough precipitation is available to keep the surface soil moist and reduce wind erosion. The shape of curve windbar depends upon the amount of rainfall and its distribution. Curve windbar would fall more steeply if the increasing precipitation is distributed more evenly or if it coincides with the summer months. Conversely, the curve would be less steep if the precipitation is mainly confined to the winter months. Unlike water erosion, the greatest opportunity for reduction of wind erosion occurs in the semi-arid range of annual precipitation — i.e. at the greatest divergence between the two wind erosion curves (windveg and windbar) °r about 200 to 500 mm. In situations where both wind and water erosion occur under a vegetation cover, there is a compensating effect between the two erosional processes, while under bare soil condition the erosional effects are additive. For example, combined water and wind erosion from vegetated surfaces (w+wveg) starts out high because of the high wind erosion potential but decrease rapidly until about 250 mm of mean annual precipitation is reached, and then reaches a secondary peak at about 400 mm of precipitation (Figure 1). With additional increases in precipitation, vegetative cover increases and water erosion decreases. In contrast, on a bare surface, combined rates of wind and water erosion (w+ Wbar) start out at high rates of erosion or sediment yield and remain high at all levels of mean annual precipitation (wind erosion essentially ceases above 950 mm of rainfall). 0 2 4 6 8 10 12 14 Mean Annual Precipitation (100 mm) 0 20% 40% 60% 80% Plant Cover Combined level of wind and water erosion from bare ground (w + wbBf) Combined level of wind and water erosion from vegetated surfaces with Increasing cover » E o o LU Cinder Noncinder Figure 2. — Levels of ectomycorrhizal colonization of pihons growing in cinder and sandy loam soils. Values are means + 1 S.E. and represent the average of four cinder and four sandy loam sites. Different letter above the bars denote significant differences at p < 0.05. Data were adapted from Gehring and Whitham (1994b). 1.25 Mutualistic Effects of Ectomycorrhizae Low High # of Ectomycorrhizae Figure 3. — Shoot weight of pihon seedlings grown in cinder soils with lower (80 or less) or higher (150 or more) numbers of ectomycorrhizal roots. Data are means +. 1 S.E. and were adapted from Gehring and Whitham (1994b). Different letters above the bars denote significant differences at p < 0.05. environments where soil fertility is low and the need for enhanced nutrient and moisture uptake is greatest (Meyer 1973). Although we do not have comparable data for the vesicular-arbuscular mycorrhizae of juniper, similar relationships between soil fertility and vesicular-arbuscular mycorrhizal colonization were found by Boerner (1986) and Van Noordwijk and Hairiah (1986). However, other researchers have observed no significant correlation between ectomycorrhizal colonization and soil fertility (e.g., Lee and Lim 1989, McAfee and Fortin 1989), and the results of studies that monitored changes in ectomycorrhizal colonization following nutrient supplementation are similarly variable, suggesting that other factors are also important (e.g., Gagnon et al. 1987, Arnebrant and Soederstroem 1990, MacFall et al. 1990, Termorshuizen and Ket 1990). Although cinder and noncinder site pinons differed in ectomycorrhizal colonization, data from seedling studies in the greenhouse demonstrate that ectomycorrhizae enhance seedling growth in both soil types. We observed a significant positive correlation between ectomycorrhizal abundance (total # of ectomycorrhizal roots) and shoot weight for seedlings grown in both cinder and noncinder soil types (Gehring and Whitham 1994b). For example, in cinder soils, the shoot weight of seedlings with high numbers of ectomycorrhizae 33 (150 or more) was approximately 44% higher than the shoot weight of seedlings with lower numbers of ectomycorrhizae (80 or less) (fig. 3). This positive relationship between ectomycorrhizae and shoot growth suggests that any environmental factor that reduces ectomycorrhizae could negatively affect pifion growth. HERBIVORES AND MISTLETOE PARASITES NEGATIVELY AFFECT MYCORRHIZAE IN PINONS AND JUNIPERS In both pifions and junipers, we observed a significant negative relationship between mycorrhizal colonization and pest attack In mature pinons, ectomycorrhizal colonization of trees resistant to moth attack was approximately 30% greater than ectomycorrhizal colonization of moth- susceptible trees in the cinder soils of Sunset Crater (Gehring and Whitham 1991). Removal of moths from susceptible trees resulted in a rebound in ectomycorrhizal colonization demonstrating that the moth was responsible for the mycorrhizal reductions in susceptible trees (fig. 4A). Younger pinons that experienced severe herbivory by a needle-feeding scale insect also suffered significant reductions in ectomycorrhizal colonization that were eliminated following insect removal (Del Vecchio et al. 1993)(fig. 4B). A similar negative relationship between pest attack and mycorrhizal colonization is found in one-seeded juniper even though junipers have a different type of mycorrhizal mutualist and are hypothesized to be better adapted to arid conditions than pinon pines. Vesicular-arbuscular mycorrhizal colonization of juniper roots was 30% lower in trees with high mistletoe densities than in trees with low mistletoe densities (Gehring and Whitham 1992) (fig. 4C). The consistent nature of these results suggests that even the mycorrhizae of more dry-adapted species will decline following pest attack To assess the importance of environmental stress to the interactions between mycorrhizae and aboveground pests, we simulated herbivory at a sandy loam site and compared the mycorrhizal responses of trees at this site to trees experiencing the same type of simulated herbivory at a cinder site. Our results indicate that environmental stress is an important component of the interaction — trees at the cinder site experienced declines in ectomycorrhizal colonization while those at the sandy loam site did not (Gehring and Whitham, in prep.). These findings suggest that environmental Ectomycorrhizae on Pinyon Pine Susceptible Susceptible Susceptible Susceptible Control Moth Control Scale Removal Removal Vesicular-arbuscular Mycorrhizae -| ■£ 80 CO ■— 2 o 60 2 ° « 75 40 3 IE 8 g 20 > o ^ E 0 C. Mistletoe Parasitism b _ a High Mistletoe Low Mistletoe Figure 4. — The relationship between moth (A) and scale (B) herbivory and ectomycorrhizal colonization in pihon pine, and the relationship between mistletoe parasitism and vesicular- arbuscular mycorrhizal colonization in juniper (C). Graphs A and B demonstrate that ectomycorrhizal levels increased following herbivore removal in pinons, while Graph C shows that vesicular-arbuscular mycorrhizal colonization of juniper roots is lower in heavily mistletoe parasitized trees. Values are means + 1 S.E. and different letters above the bars denote significant differences at p < 0.05. Data were adapted from Gehring and Whitham 1991, 1992, and Del Vecchio et al., 1993). stress not only influences the responses of mycorrhizae and herbivores individually, but also influences the interactions between the two. Trees growing in more benign environments may be able to retain their mycorrhizae following herbivory or other pest attacks while trees growing in more stressful environments cannot. MANAGEMENT IMPLICATIONS Our findings, summarized in Table 2, have several potential implications for the management 34 Table 2 - Summary of interactions among environmental stress, pest attack, and mycorrhizal colonization in pinyons and junipers. Higher Environmental Stress Lower Environmental Stress Factor Pinyon Juniper Pinyon pest attack mycorrhizae high moth herbivory high ectomycorrhizal colonization mycorrhizae-pest natural or simulated interaction herbivory causes declines in ectomycorrhizae high mistletoe parasitism (9 > cT ) no data high levels of mistletoe parasitism associated with declines in veslcular-arbuscular mycorrtiizae low moth herbivory low ectomycormizal colonization no effect of simulated herbivory on ectomycorrhizae Juniper low mistletoe parasitism ( g = c ) no data no data of pinon-juniper woodlands. First, our results suggest that pifions and junipers are more likely to experience attack by three different pests as environmental stress increases. This finding is especially noteworthy because these pests exhibit the same patterns even though they have different lifestyles and feeding modes; a parasitic mistletoe that feeds from the xylem sap, a scale insect that feeds from the mesophyll of needles from juvenile plants, and a stem- and cone-boring moth that attacks mature trees. These results have implications for broad issues associated with global climate change. For example, if environmental threats such as global warming cause soil in more benign environments to become warmer and drier (two characteristics of the stressful environments we studied), our findings suggest that trees growing in warmer conditions would experience increased susceptibility to attack by a variety of pests and would, as a result, suffer reduced rates of growth and reproduction. Ayres' (1993) observation that the rate of increase of moth populations on mountain birch trees was 2.9-fold higher on trees growing in a greenhouse warmed by only one degree Celsius relative to controls supports this hypothesis. Ayres (1993) predicted that global warming of only 2 to 4 degrees would lead to outbreaks of many insect herbivores if insects are generally more sensitive to temperature increases than their host plants. However, we believe that plant-herbivore responses to global warming are likely to be more complex than this. For example, 75% of the 450 studies reviewed by Waring and Cobb (1992) reported significant herbivore responses to plant stress. These responses varied from positive to negative to nonlinear and depended upon the types of plants (e.g., flowering plants vs. conifers) and the resource requirements of the herbivores involved (Waring and Cobb 1992). Second, our data suggest that the mycorrhizae of pinon-juniper woodlands are also likely to be affected by global climate change, although the response of mycorrhizae is less clear. Our finding that pinons growing in stressful cinder environments have higher ectomycorrhizal levels than pinons growing in sandy loam soils leads us to predict that mycorrhizal colonization would increase with global warming. We found that pinons growing in sandy loam soils had 33% higher levels of ectomycorrhizal colonization during a drier year than a moister year, supporting this hypothesis (Gehring and Whitham, unpub data). However, we have shown that pests such as herbivores and mistletoes have a negative effect on mycorrhizae and have hypothesized that the densities of these pests were likely to increase with global warming. Therefore, any increases in mycorrhizae observed in response to the increased abiotic stress of warmer, drier soil could be offset by decreases in mycorrhizal colonization in response to increased pest attack The mycorrhizae of some herbaceous species, especially grasses, have been shown to be reduced by herbivory (e.g., Bethlenfalvay et al, 1988, reviewed by Gehring and Whitham 1994a) suggesting that pinon-juniper community understory species may exhibit similar complicated relationships with environmental stress and parasite attack. Third, the negative association we observed between aboveground pests and belowground mutualists emphasizes the importance of studying both components of the plant (shoot and root systems) in concert. Assessment of damage to aboveground tissues alone does not give us a complete understanding of the impacts of aboveground pest attack. For example, aboveground herbivory has negative effects on pirion ectomycorrhizae in cinder soils, but not in 35 sandy loam soils, suggesting that the effects of aboveground herbivory on belowground processes can vary with environmental stress and may not be easily predicted. The same is likely to be true for other aboveground stressors such as grazing, some fires, etc. Finally, the patterns of increasing herbivory and mycorrhizal reductions on trees growing in cinder soils have additional management implications because they are associated with host plant genetics. Mopper at al. (1991) found genetic differences both between pihons growing in cinder and noncinder soils, and between moth-resistant and moth-susceptible pihons growing in the cinders. Furthermore, Cobb and Whitham (1994) observed an association between growth and genotype in pinons growing in the cinders, and found that the slowest growing genotypes were significantly less abundant in mature trees than juvenile trees suggesting that selection had favored the fastest growing genotype. These differences suggest that the abilities of pihons at Sunset Crater to survive and to resist herbivory are likely to be genetically based. These results have two additional implications for the management of pinon-juniper woodlands. First, these genetic data combined with data on the associations between pihon community members and pifion insect resistance, argue that there is a genetic basis to pihon community structure that links community members. We have already described linkages between mycorrhizal mutualists and insect-resistant and insect- susceptible pihons. In addition, Christensen and Whitham (1991, 1993) found that moth herbivory on insect-susceptible trees had significant impacts on the birds and mammals dependent upon pihon seeds for food. Thus, through its negative effect on the cone production of genetically susceptible trees, moth herbivory caused a shift in the balance between birds and mammals. At stressed sites with high levels of moth herbivory, the majority of the pihon crop was harvested by mammals, who act as seed predators, rather than by birds who are important seed dispersers (Christensen and Whitham 1991, 1993). Such plant genetics-based linkages between diverse taxa in pihon-juniper woodlands may make this community more susceptible to environmental perturbations than other communities where members are less tightly linked or are loose assemblages of species. Therefore, management decisions that affect one community member (e.g., pihon pine) may have repercussions on the entire community dependent upon that species. Second, because plant genetics plays a role in pihon drought tolerance and resistance to insect herbivory, environmentally stressful areas such as Sunset Crater are important to conserve because of the unique plant genotypes they possess. For example, Sunset Crater and other cinder sites may contain some of the most drought-tolerant and insect-resistant pihon genotypes. These stress- tolerant genotypes could be commercially important for outplanting in arid areas, and could serve as adapted gene pools in the event of global warming and pihon die-back. Thus, although it is important to preserve the center of a species distribution, where its growth and reproduction is greatest, the conservation of more extreme boundary populations under greater biotic and abiotic stress may be of equal importance. ACKNOWLEDG EM ENTS The work reviewed here was supported by USDA grants 91-37302-6224 and 92-37302-7854, DOE grant DE-FG03-94ER61849, and NSF grant BSR-9107042. The authors thank the U.S. Forest Service for access to the study sites. LITERATURE CITED Arnebrant, K., Soederstroem, B. 1990. The influence of nitrogen fertilization on ecotmycorrhizal mycelial colonization and infection. Agriculture Ecosystems and Environment 28: 21-25. Ayres, M.E 1993. Plant defense, herbivory, and climate change. In: Biotic Interactions and Global Change. EM. Kareiva, J.G. Kingsolver, and R.B. Huey, eds., Sinauer Associates Inc., Sunderland, MA, pp. 75-94. Bethlenfalvay, G.J., Thomas, R.S., Dakessian, S., Brown, M.S., Ames, R.N. 1988. Mycorrhizae in stressed environments: effects on growth, endophyte development, soil stability, and soil water. In: Arid Lands Today and Tomorrow. E.A. Whitehead, C.F. Hutchinson, B.N. Timmerman, and R. G. Varady, eds., Westview Press, Boulder, CO, pp. 1015-1029. Boerner, R.E.J. 1986. Seasonal nutrient dynamics, nutrient resorption, and mycorrhizal infection intensity of two perennial forest herbs. American Journal of Botany 73: 1249-1257. Christensen, K.M., Whitham, T.G. 1993. Herbivore impact on competition between birds and mammals for pinyon seeds. Ecology 74: 2270-2278. Christensen, K.M., Whitham, T.G. 1991. Indirect herbivore mediation of avian seed dispersal in pinyon pine. Ecology 72: 534-542. 36 Cobb, N.S., Whitham, T.G. 1994. Genetic variation associated with chronic water and nutrient stress in pinyon pine. American Journal of Botany 81: 936- 940. Cobb, N.S., Mopper, S., Gehring, C, Caouette, M., Christensen, K.M., Whitham, T.G. Herbivore response to stress and experimental stress release in pinyon pine. Submitted to Ecology. Del Vecchio, T., Gehring, C.A., Cobb, N.S., Whitham, T.G. 1993. Negative effects of scale insect (Matsucoccus acalyptus) herbivory on the ectomycorrhizae of pinyon pine. Ecology 74: 2297-2302. Ehleringer, J.R., Schulze, E-D, Ziegler, H., Lange, O.L., Farquhar, G.D., Cowan, I.R. 1985. Xylem-tapping mistletoes: water or nutrient parasites? Science 227: 1479-1481. Ferrel, G.T 1974. Moisture stress and fir engraver (Coleoptera: Scolytidae) attack in white fir infected by true mistletoe. The Canadian Entomologist 106: 315-318. Freeman, D.C., McArthur, E.D. 1982. A comparison of twig water stress between males and females of six species of desert shrubs. Forest Science 28: 304-308. Freeman, D.C., Klikoff, L.G., Harper, K.T. 1976. Differential resource utilization by the sexes of dioecious plants. Science 193: 597-599. Gagnon, J., Langois, C.G., Fortin, J.A. 1987. Growth of containerized jack pine seedlings inoculated with different ectomycorrhizal fungi under a controlled fertilization schedule. Canadian Journal of Forest Research 17: 58-62. Gehring, C.A., Whitham, T.G. 1994a. Interactions between aboveground herbivores and the mycorrhizal mutualists of plants. Trends in Ecology and Evolution 9: 251-254. Gehring, C.A., Whitham, T.G. 1994b. Comparisons of ectomycorrhizae on pinyons pines (Pinus edulis; Pinaceae) across extremes of soil type and herbivory. American Journal of Botany, in press. Gehring, C.A., Whitham, T.G. 1992. Reduced mycorrhizae on Juniperus monosperma with mistletoe: the influence of environmental stress and tree gender on a plant parasite and a plant-fungal mutualism. Oecologia 89: 298-303. Gehring, C.A., Whitham, T.G. 1991. Herbivore-driven mycorrhizal mutualism in insect-susceptible pinyon pine. Nature 353: 556-557. Hikmat, A.A.A., Strain, B.R., Mooney, H.A. 1972. The physiological ecology of diverse populations of the desert shrub, Simmondsia chinensis. Journal of Ecology 60: 41-57. Hull, R. J., Leonard, O.A. 1964. Physiological aspects of parasitism in mistletoes (Arceuthobium and Phoradendron) I. The carbohydrate nutrition of mistletoes. Plant Physiology 39: 996-1007. Knutson, D.M. 1983. Physiology of mistletoe parasitism and disease responses in the host. In: The Biology of Mistletoes. D. Calder, and P Bernhardt, eds., Academic Press, Sydney, pp. 295-316. Lee, S.S., Lim, K.L. 1989. Mycorrhizal infection and foliar phosphorous content of seedlings of three dipterocarp species growing in a selectively logged forest and a forest plantation. Plant and Soil 117: 237- 241. MacFall, J.S., Iyer, J., Slack, S., Berbee, J. 1990. Mycorrhizal-phosphorous interaction on red pine (Pinus resinosa Ait). Agriculture Ecosystems and Environment 28: 321-324. Marshall, J.D., Ehleringer, J.R. 1990. Are xylem-tapping mistletoes partially heterotrophic? Oecologia 84: 244- 248. McAfee, B.J., Fortin J.A. 1989. Ectomycorrhizal colonization on black spruce and jack pine seedlings outplanted in reforestation sites. Plant and Soil 116: 9- 17. Meyer, F.H. 1973. Distribution of ectomycorrhizae in native and man-made forests. In: Ectomycorrhizae: Their Ecology and Physiology. J.C. Marks and T T. Kozlowski, eds., Academic Press, New York, NY, pp. 79-106. Mopper, S., Mitton, J., Whitham, T.G., Cobb, N., Christensen, K.M. 1991. Genetic differentiation and heterozygosity in pinyon pine associated with herbivory and environmental stress. Evolution 45: 989-999. Stark, N. 1970. Water balance of some warm desert plants in a wet year. Journal of Hydrology 10: 113- 126. Termorshuizen, A.J., Ket, PC. 1990. The effects of fertilization with ammonium and nitrate on mycorrhizal seedlings of Pinus sylvestris. Agriculture Ecosystems and Environment 28: 497-501. Van Noordwijk, M., Hairiah, K. 1986. Mycorrhizal infection in relation to soil pH and soil phosphorous content in a rain forest of northern Sumatra. Plant and Soil 96: 299-302. Waring, G.L., Cobb, N.S. 1992. The impact of plant stress on herbivore population dynamics. In: Plant-Insect Interactions., E. A. Bernays, ed., CRC Press, Boca Raton, FL, pp 167-226. White, T.C.R. 1984. The abundance of invertebrate herbivores in relation to the availability of nitrogen in stressed food plants. Oecologia 63: 90-105. White, T.C.R. 1976. Weather, food, and plagues of locusts. Oecologia 22: 119-134. White, T.C.R. 1974. A hypothesis to explain outbreaks of looper caterpillars with special reference to populations of Selidosema suavis in a plantation of Pinus radiata in New Zealand. Oecologia 16: 279-301. White, T.C.R. 1969. An index to measure weather- induced stress of trees associated with outbreaks of psyllids in Australia. Ecology 50: 905-909. Whitham, TG., Mopper, S. 1985. Chronic herbivory: impacts on architecture and sex expression of pinyon pine. Science 227: 1089-1091. 37 Stand Dynamics on Upper Elevation Pinon- Juniper Watersheds at Beaver Creek, Arizona Gerald J. Gottfried1 and Peter F. Ffolliott2 Abstract. — There is a lack of information about stand dynamics, especially volume growth, in the pinon-juniper woodlands of the southwestern United States. Such information is vital for managing woodlands on a sustainable basis for tree products. Harvesting in excess of growth will diminish the resource. Growth information is also needed to understand ecosystem dynamics and to ascertain stand changes that affect other resources, such as wildlife habitat. Two overstory inventories, spanning a 24-year period, were conducted on permanent points on two untreated watersheds in central Arizona. Alligator juniper (Juniperus deppeana) is the dominant species on the watersheds, which are representative of woodlands at the upper elevations along the Mogollon Rim. Although inventory procedures have changed the first inventory in 1964, the repeated measurements utilizing the original techniques can produce valuable information about stand changes over the intervening period. Periodic annual growth for trees present in 1964 and 1988 was 18.5 cubic feet per acre and net periodic annual growth was 20.3 cubic feet per acre. Relationships between diameter at breast height (dbh) class and equivalent diameter at root collar (edrc) and between edrc and total height or crown area were developed for alligator juniper. INTRODUCTION Healthy and sustainable ecosystems are goals of ecosystem management. Information on stand dynamics, particularly growth, is essential for the management of any ecosystem. Excessive harvest- ing or mortality, above growth levels, will diminish the resource, compromising sustainability objec- tives. Basic stand growth and dynamics informa- tion is also necessary in order to understand ecological processes within an ecosystem and can be used for planning and evaluating impacts of a variety of management treatments. Information on tree growth, mortality, and stand dynamics in commercial forests and in many woodlands is rou- tinely collected by managers and researchers. Some studies have followed stand dynamics for decades. 1 Research Forester, USD A Forest Service, Rocky Mountain Forest and Range Experiment Station, Flagstaff, AZ. Headquarters is in Fort Collins, CO. 2 Professor, School of Renewable Natural Resources, University of Arizona, Tucson, AZ. Although pinon-juniper woodlands cover large ar- eas of the Southwest, similar data collections have only begun in recent years. One reason for the de- lay is that ecosystem management of pinon-juniper woodlands for a number of resources and ameni- ties is a relatively recent development (Gottfried and Severson 1993). Following World War II, trees were eradicated from large areas in the hope of improving forage for livestock production, water yields, and wildlife habitat. The value of destroyed tree resources was not considered. Another reason for the lack of multiresource management is diffi- culty of obtaining data. In most temperate forests, past growth can be determined by increment core samples from trees. This procedure will work in woodlands where pirion is the main species or an important component. However, increment cores 38 have less value in stands where junipers dominate, since these trees have many false and missing rings. In the late 1950s and 1960s, a watershed evaluation program was established in central Ari- zona to test hypotheses concerning the relation- ships between vegetation manipulations and the augmentation of water yields. The program was concerned with the ponderosa pine (Pinus ponder- osa) and pinon-juniper vegetation types. Research efforts also included evaluating the effects of a se- ries of treatments on livestock forage, timber pro- duction, wildlife habitat, recreational values, and erosional dynamics. As part of the effort, repeated overstory inventories were conducted at perma- nent inventory points to ascertain changes result- ing from treatments. Some watersheds, designated as control areas, were not treated. Two pinon-juniper watersheds, dominated by alligator juniper (Juniperus deppeana), were not treated and, therefore, provide the oppor- tunity to evaluate stand changes from 1964 to 1988. Woodland inventory techniques had not been re- fined in 1964 when the watersheds were first in- ventoried. The inventory crew adapted procedures commonly used in ponderosa pine forests to juni- per species characterized by multi-stemmed growth forms. Although woodland inventory pro- cedures have changed (Meeuwig and Budy 1981), repeated measurements using the original proce- dures provide valuable information about stand dynamics over the 24-year period. The information, although from one case study, could be applicable to similar alligator juniper stands within the Southwest. STUDY AREA The Beaver Creek Study Area is located south of Flagstaff along the Mogollon Rim and within the Verde River Basin of central Arizona. The area within the Coconino National Forest; the USDA Forest Service is responsible for administration of activities not related to research. Elevations in the watersheds range from 3,000 to 8,000 feet. Beaver Creek contains 18 small experimental watersheds, six of which support pinon-juniper stands. These stands are located on sloping mesas and breaks, steep canyons, and valleys; elevations range from 4,500 feet to 6,500 feet (Clary et al. 1974). Utah juni- per (J. osteosperma) is the dominant tree species on three of the watersheds; alligator juniper, the spe- cies of interest, is dominant on the others. Pinon-juniper woodlands are generally located at 7,500 feet in the Southwest. The alligator juniper stands at Beaver Creek are present at elevations less than 7,500 feet but are considered upper ele- vation stands in areas along the Mogollon Rim. Al- ligator juniper, a representative of the Madrean flora of Mexico, is present above 6,000 feet in Ari- zona and in the southern two-thirds of New Mex- ico (Gottfried and Severson 1993). It is also the most common juniper in northern Mexico. Repeated inventories of the two untreated alli- gator juniper watersheds are the basis for the study. These areas are designated Watershed No. 4 and Watershed No. 5. The characteristics of these watersheds have been presented by Clary et al. (1974), Baker (1982), and Pollisco (1987). Average elevations of these watersheds are 6,200 to 6,400 feet, where annual precipitation averages about 20 inches and ranges from 16 to 27 inches (Clary et al. 1974). Approximately 62 percent of the precipita- tion falls from October through April (Baker 1982). Mean annual temperature is 50 F. The soils on the two watersheds primarily be- long to the Springerville series, with Springerville very stony clay, 0 to 10 percent slopes being most common (Williams and Anderson 1967). Springerville soils, which developed from weath- ered basalt and cinder materials, are classified as Typic Chromusterts. Basalts and cobblestones can cover from 30 to 50 percent of the surface. The tree species on the two areas included alli- gator juniper, Utah juniper, pinon, ponderosa pine, and Gambel oak (Quercus gambelii). The taxonomy of the pinon is still unclear. It generally has one needle in a fascicle and has been classified as Pinus edulis var. fallax, P. fallax or P. californiarum subsp. fallax. The two watersheds had similar basal areas, but the distributions of the species were different (table 1.) Alligator juniper was the most common species on both watersheds, with 55 and 86 percent of the basal areas, respectively. The high number of alligator juniper trees on Watershed 5 is because of the larger number of young trees. No effort was made to separate trees of sprout or seed origin. Utah juniper and ponderosa pine appeared to be more common on Watershed 4. Ponderosa pine was present along the drainages and on the upper slopes. Alligator juniper constitutes 69 percent of the basal area on the combined stand and 75 per- cent of the trees (table 1). Blue grama (Bouteloua gracilis) is the principal understory species. The vegetation has been classified as belonging to the Juniperus deppeana/Bouteloua gracilis Habitat Type (USDA Forest Service 1987). 39 Table 1. — Stand conditions in 1964 on Beaver Creek Watershed 4, Watershed 5, and the combined watersheds. (Tree and basal area (square feet) values are on a per acre basis.) Combined Watershed 4 Watershed 5 Watersheds Area (acres) 333 64 397 No. inventory points 63 46 1 09 Species Trees Basal Area Trees Basal Area Trees Basal Area Alligator juniper 95.2 22.6 471.3 37.0 253.9 28.7 Pinon 0.0 0.0 31.8 1.6 13.4 0.7 Utah juniper 48.3 7.5 56.0 1.6 51.6 5.1 Ponderosa pine 33.6 9.9 1.0 2.2 19.8 6.7 Gambel oak 1.9 0.8 0.7 0.5 1.4 0.7 Total 179.0 40.8 560.8 42.9 340.1 41.9 METHODS Information on tree growth, mortality, and stand dynamics was based on the repeated meas- urements of trees surrounding permanent inven- tory points in 1964 and 1988. The point locations were determined according to a multiple random start design (Shiue 1960). The points were on tran- sects established perpendicular to the main stream channel on each area and extended to the water- shed boundaries. Tree information was collected at every third point along the transects. The distances between points and, thus, the number of points were different on the areas (table 1). Inventory data were combined to improve representation of con- ditions in pinon-juniper stands characterized by alligator juniper. All measurements were made on trees selected by variable plot (point) sampling based on a 25 BAF angle gage. A total of 109 points were included in the analysis; several points were eliminated be- cause of questionable data. The 1964 inventory adapted techniques commonly used in ponderosa pine forests to the woodlands. In this inventory, the gage was aimed at the breast height (bh) to deter- mine if a tree was to be included in the sample, and the species and diameter at breast height (dbh) were determined. Trees that were forked below bh were tallied as separate trees. The measurement lo- cations were not permanently marked on the trees. Current procedures for using point sampling in pi- non-juniper woodlands call for the angle gage to be sighted at stump height or at the root collar (rc) (Meeuwig and Budy 1981). Fixed area plots are used by some woodland managers and researchers (Chojnacky 1988). In addition, woodland tree di- ameter measurements are commonly made at rc not bh. However, the use of dbh is easier to justify with alligator juniper, since older trees are often single-stemmed. To replicate the 1964 inventory, the crews used the same procedures in 1988. Bor- der trees were checked with a tape to compare tree distance from the point to the critical distance based on dbh and the BAF. In 1988, measurements were also made at drc, and measurements of total height and mean crown diameter were included. Mean crown diameter is the geometric mean of the longest crown diameter and the perpendicular di- ameter, and represents the vertical crown projec- tion. Only dbh and total height were measured on ponderosa pine and Gambel oak Tree regeneration was not sampled. The 1964 data were recorded by 2-inch diame- ter class; for example, the 10-inch class includes trees with diameters between 9.0 and 10.9 inches. In 1988, measurements of drc and dbh were made to the nearest 0.1 inch. Equivalent dbh (edbh) and equivalent drc (edrc) were calculated for stems over 1.5 inches on multi-stemmed trees (Chojnacky 1988). Growth was based on changes in dbh class over the period; this is appropriate because bh was not marked on sample trees. Some trees showed declines in dbh over the period. Since past growth cannot be determined from increment cores, these trees were assigned the 1988 diameter, providing a more conservative estimate of growth. Growth was calculated using standard proce- dures (Husch et al. 1972, Chambers 1984) and fol- lowing the example described by Gottfried (1992) for a southwestern mixed conifer stand. Changes were based on trees present in 1964 and 1988; these are classified as "survivor trees". The 1988 trees per acre values include ingrowth into the 2-inch class, while ingrowth for volume includes new trees that grew into the 2-inch class or survivor trees that grew into the 6-inch class for ponderosa pine. On- growth trees, other than those qualifying as in- growth, are not included; these are trees of any size that were too small to be measured in 1964 but have grown sufficiently during the interim to be included in the subsequent inventory. In variable plot sampling, ongrowth does not represent past periodic growth but serves as a basis for future growth evaluations. 40 Volume for pirion and the junipers was based on the cubic foot volume equations developed by Chojnacky (1988) for Arizona woodland species. Regression relationships were developed between the 1988 dbh class and edrc (figure 1) to make the volume calculations. A similar relationship was cal- culated by Chojnacky (1988) using data from the San Carlos Apache Reservation in south-central Arizona (figure 1). A relationship between edrc and total height also was required (figure 2). Chojnacky (1988) only used data from single-stem trees while data from single and multi-stem trees were com- bined for the development of the Beaver Creek re- lationships. Approximately 20 percent of the junipers in the Beaver Creek sample had multiple stems. Ponderosa pine volumes were based on equations developed by Myers (1963) and oak vol- umes were based on tables presented by Barger and Ffolliott (1972). Regression relationships were developed be- tween edbh and edrc, edrc and total height, and edrc and crown area since there is a lack of infor- mation about characteristics of alligator juniper. Several linear and multiple regression models were evaluated, and the best, based on the coefficients of determination (r ), are presented. 35 30 25 0 ■4—" sz O) 20 t the other chainlngs This 1 1 ho\N t\ W ttu* only site whore squirreltail is not in the most preferred group Alteration ot \ ogotatton composition by the presence ot tree competition! such as the reduction ot larger bunchgrasses, may be responsible U>i this difference Overall, the absence ot trees on this site has not affected the general vegetation preferences ot mule door Deer Utilization Summary All the sites studied were located in known deer use areas and were undertaken in .in attempt to improve doer habitat in those areas In the clus ter analysis results deer use was closely associated with the same general vegetation complex ovor all the ehaiuiugs. This indicated a possible rotation ship between the abundance ot the preferred plant species on a site and the level ot deer use that occurred following chaining I ho average percent vovor ot the plant spoeios in the preferred group identified by elnster analysis was determined tor each site This was compared with the index ot average deer-days use per acre on each site using regression analysis (big. o). For the five chaiuings studied, cover ot the preferred native plant species i o V) z> O ^0 ■ M N 20 15 10 0 • ra=0.76 L 1 - L 0 5 10 15 Cover of Preferred Vegetative Group Ptgura t. — Ra^r««*kw analysis of tha parcant cover of tha plant •pacta* In tha prafarrad group* In figure* i - 5 and tna dear day a uaa par a era ^POA) on flva chaining*. CONCLUSIONS General patterns were evident in the results. Areas ol dosed stand woodlands with older trees Consistently had the lowest deer use after treat ment. Ihis occurred even when the chained area was seeded. The seeded exotic grasses were the |]\OSt Successful On these locations hut apparently not preferred by mule deer. Native species that were seeded as part ot the treatment did poorly on the tormerly tree dominated sites (Tausch lausch and ' Inciter N77). Mvcorrhi/al associations in the soil are a potentially important factor in the plant establishment patterns observed following treatment Pifton has different associations than juniper and the native uuderstorv species. The Introduced annuals such as russian thistle and tumble mustard and the exotic hunch grasses seeded on the site do not require mvcorrhizal associations (Klopetek et al< 1969). Ultimately these differences in mycorrhiia] associations affect vege- tation patterns which in turn affect deer use. Better WaVS tor returning native understory species to tree dominated sites following treatment need to he developed. When Sufficient uuderstorv was present prior to chaining it responded vigorously to the treat- ment. Not all such uuderstorv communities, how- ever were the most favored In mule deer. Selectiv- ity was also occurring among these communities Those most preferred were generally sites with i good mix of shrubs, grasses and torhs. In another part ot this study published else- where (lausch and Uieller lv>7T) sucoessioual pat- terns indicated that the benefits ol the chaining procedure tor deer weie short lived. Accelerated growth of surviving and reestablished trees on the ehainings studied was rapidly reducing the in- creased uuderstorv production. All ot them were projected to return to pre-chaiuing levels of pro- duction and deer use in less than 20 years. While trees provide horizontal and vertical thermal cover (Sumiuski 1993), continuous stands have been observed throughout the southwest to result in a reduced torage base in the uuderstorv (Gottfried and Severson W°3). However total removal ot the trees resulting in large areas of only the former uuderstorv communities, is also not effective (Severson and Medina 1983, Tausch 1973), Periodic treatment of the preferred areas with Sufficient uuderstorv present, but while leaving areas of trees close bv should be a wav to maintain useful productivity for deer w inter range (Short et al. 1977). A number of alternative methods of tree control such as burning may represent better 72 ways to accomplish this for many anas However, all methods can have other associated negative ecosystem effect*. LITERATURE CITED Berg, R.H. 1966. An evaluation of selected Nevada deer range',: f ondition, forage potential and deer livestock competition. MS 'Ihesis. University of Ne- vada, Reno, NV. Cole, N.J. 1968. Mule deer utilization of rehabilitated Nevada rangelands. M.S. Theftis. University of Ne vada, Reno, NV. Gottfried, f,.J , and Severson, K.P. 1993. f distribution and Multiresource Management of pin yon juniper woodlands in the Southwestern United States /«: Symposium: Managing pin yon juniper ecosystem for sustainability and social needs. J993, April 26 30, Santa Fe, New Mexko. Port Cbffiflf, CO: U.S. De- partment of Agriculture, Forest Service, Rocky Moun tain f orest arid Range P.xperimenf Station 108 1 16 Humphreys, M. fed.;. 1966 Big game range inventory, Phase P P/tent of seasonal big game range maps. Federal aid project W-43-R-1 No. 4. Wopatek, CC, Debano, L.F., and KlopateP, J.M. 1988 F.ffects of simulated fire on vesicular arbuscular my- corrhizae in pin yon -juniper woodland soil Plant and Soil 109:245-249 Ludwig, J.A., and Reynolds, J. P. 1990. Statistical F/ology. John Wiley and Sons, New York Neff, DJ. 1968. The pellet group count technique for big game trend and distribution: A review journal Wild life Management 32.7)7 614 Severson, K.F.., Medina, A. P. 1983 fdeer and PIP habitat management in the Southwest, Journal of Range Management Monograph No. 2. Short, IIP, Pvans, W., and BoeFer, IP 1977 The „ -.«• of natural and modified pinyon pine woodlands by deer and elk Journalof Wildlife Management H-.rA&TM). Short, f I P, and Mcf iilloc h, f / 1977. Managing pinyon juniper range-. for wildlife c j«., |, Pf/I 47 Fort r ollins r r >> U.S. I department of Agru ultun-, I or <• ,t Service, Rocky Mountain Forest and Range P.x p<-nm«-n» Station SoFal, R.R., and Sneath, I'M 1963 Princ iples of numeri ca I taxonomy. W.H. Freeman and r o San f ranc isco Suminsh, P P J98S Mule deer managefnent an a 23. Unpublished report on file at Ply PPM l/i;.trict Office. Fly, NV. Suminslo, P P. 1993. Management implication*; for mule deer Winter range in Northern Pinyon juniper. /«; Symposium Managing pinyon juniper «'o-./.u-m for sustainability and social needs 1993, April 7/, 30, Santa Fe, New Mexico. Fort r ollins, r O: US fx- partmenf of Ag/ic ultun-, Forest Se rvice, Rocjcy Moun tain Forest and Range Pxperiment Station 133 139 fausch, R.J. 1973. Plant suc/*-ssion and mule- deer utiliza tion on pinyon juniper chaining;, in Mevada. MS Thesis, University of Nevada, Reno, NV. fausch, P j , and fueller, PI 1977 Plant suc/ession fol lowing chaining of pinyon juniper woodlands if. eastern Nevada. Journal of Range Management 30 44 49. Tausch, R.J., West, N.E., and Nabi, A .A. I98P free age and dominance patterns in the Oreat Basin pinyon juniper woodlands. Journal of Range Management 342,9 2M West, N,E. 1984 Factors affecting treatment suc/ess in the pinyon juniper type. /«: Jc, hn son, P F fed ^ f'ro ceedings, Second Sfuub f/ology 7/orlo.hop, Utah State Pnr/er sity, f/>gan, LT1 pp 21 23. West, N.E., Tausch, R.J., Rea, K.f P, and fueller, ff 1978 Phytogeographical variation within juniper pinyon wood la nds of the '/real Basin r/re,,t Basin Naturalist Memoirs 2:119 136. 73 Characteristics of Pinon- Juniper Habitats Selected for Feeding by Wintering Merriam's Turkey Brian F. Wakeling and Timothy D. Rogers1 Abstract. — We studied winter habitat selection by Merriam's turkeys (Meleagris gallopavo merriami) to determine characteristics of pinon (Pinus edulis) -juniper (Juniperus spp.) habitats that may be critical to their overwin- ter survival. Seventy radio-instrumented turkeys were monitored during 4 years on the Chevelon Study Area in northcentral Arizona. Habitat charac- teristics were measured at 22 feeding sites and 17 random plots within the pinon-juniper cover type and the ponderosa pine (Pinus ponderosa)-p\f\or\- juniper ecotone. Gambel oak (Quercus gambelii) was more prevalent at feeding sites than at random plots. Feeding sites were located in the prox- imity of smaller canopy openings more frequently than random plots. Forb cover was greater at feeding sites than random plots. Management strate- gies that 1) protect and encourage mature mast producing species, espe- cially Gambel oak, and 2) preserve small (<0.06 ac) openings <1 mi from suitable roosting habitat will favor turkey populations. INTRODUCTION Merriam's turkey seasonally inhabit pinon- juniper forest cover type (Shaw and Mollohan 1992). Although this cover type comprises only a small portion of their annual use areas (Wakeling 1991), turkeys rely on pinon-juniper mast during severe winters (Ligon 1946). Winter food sources, such as pinon seeds and juniper berries (Ligon 1946, Reeves 1953, Schorger 1966), are more abun- dant in this cover type. Because of these character- istics, pinon-juniper habitats have been described as ideal emergency winter range (Reeves 1954). Land management practices in the pinon- juniper woodland alter characteristics of the habi- tat used by turkeys. Turkey habitat use is influ- enced by grazing, fuel wood harvest, timber treat- ments, and recreational activities (Ligon 1946, Schorger 1966, Scott and Boeker 1977, Shaw 1986, Hoffman et al. 1993). Our research was designed to identify characteristics of pinon-juniper forest cover types selected by Merriam's turkey because little quantitative information was available to 1 Brian Wakeling is a Research Biologist and Tim Rogers is a Re- search Assistant with the Arizona Game and Fish Department, Phoenix, AZ. assist in land management decisions. We tested null hypotheses that turkeys used characteristics of feeding habitat proportionate to their measured availability. Funding for this project was provided through Federal Aid in Wildlife Restoration Act W-78-R. The U. S. Forest Service Chevelon Ranger District of the Apache-Sitgreaves National Forests provided logistical assistance. J. S. Elliott, C. H. Lewis, J. Sacco, K. Sergent, and C. A. Staab provided field assistance. S. G. Woods assisted with GIS expertise. We would like to thank S. S. Rosenstock and R. A. Ockenfels for reviewing earlier drafts of this manuscript. STUDY AREA The 335 mi2 Chevelon Study Area (CSA) was lo- cated on the Mogollon Rim, approximately 40 mi south of Winslow, Arizona, on the extreme western edge of the Apache-Sitgreaves National Forests (Fig. 1). Elevations ranged from 5500 ft in the northern portion to 7900 ft in the southern portion. 74 Annual precipitation averaged 18.6 in, with 2 con- centrations. The first concentration occurred dur- ing winter storms in January through March, and the second during summer monsoon storms in July through early September (Natl. Oceanic and At- mos. Admin. 1991). Five cover types were identified on the CSA based upon U. S. Forest Service Terrestrial Ecosys- tem Surveys (Laing et al. 1989): 1) pinon-juniper, 2) ponderosa pine-Gambel oak, 3) mixed conifer, 4) aspen (Populus tremuloides), and 5) forest meadow cover types (Fig. 2). Mixed conifer cover types were dominant above 7600 ft, and extended along east facing slopes and drainages. This habitat included Douglas-fir (Pseudotsuga menziesii), white fir (Abies concolor), limber pine (Pinus flexilis), and Rocky Mountain maple (Acer glabrum). Ponderosa pine dominated west facing slopes between 7600 and 6000 ft. Below 6000 ft, the pihon-juniper cover type was dominant, with ponderosa pine stringers along drainages. At elevations below 7000 ft, pinon pine and alligator juniper (Juniper us deppeana) were increasingly abundant. Gambel oak occurred in all wooded cover types, in pockets in the mixed coni- fer and pinon-juniper associations, and as a wide- spread codominant with ponderosa pine. Figure 1.— Location of the Chevelon Study Area in northcentral Arizona. Logging and grazing were the major commer- cial land uses on the CSA. Cutting of fuel wood, particularly in the pinon-juniper cover type, has increased over the past 2 decades. Logging began in the late 1930's and most ponderosa pine stands on level terrain have been logged at least once. However, little logging has occurred on steeper slopes in larger canyons. Until the 1960's, sheep were the primary livestock on the CSA. The pre- dominant livestock on the CSA since the 1960's has been summering cattle. METHODS Merriam's turkey were captured between 1 January and 31 March during the winters of 1988- 92. Turkeys were captured with box traps, drop nets, and rocket nets (Glazener et al. 1964, Bailey et al. 1980, Phillips 1982, Wakeling 1991). Each turkey was fitted with a backpack mounted radio teleme- try unit (Telonics, Mesa, AZ) (Wakeling 1991). These birds were observed to determine habitat selection between 15 November and 15 April during the winters of 1990-91 through 1993-94. We visually located radio-instrumented turkeys or feeding sign (e. g. scratching and droppings) from instrumented or non-instrumented turkeys to determine the activity center. We used this point as plot center in the mensuration of habitat character- istics. Locations were obtained approximately 2 times daily. Habitat mensuration was conducted within 2 days of when the birds abandoned the feeding site. Individual turkeys were not located >1 time per day to reduce autocorrelation of data. We classified vegetative cover types at the sites according to Larson and Moir (1986). We consid- ered the following Larson and Moir (1986) classifi- cations to represent pinon-juniper habitats for our analysis: pihon pine-blue grama (Bouteloua gracilis), pihon pine-sparse, pinon pine-cliffrose (Cowania mexicana), and the pinon pine phase of ponderosa pine-Gambel oak. A 0.1-ac circular plot was used to estimate density by counting conifer and Gambel oak seedling (<1 in diameter at breast height [DBH]) and trees (>1 in DBH). We measured the DBH of all ponderosa pine and Gambel oak trees on the 0.1-ac plot with a diameter tape. The diameter at root crown (DRC) was measured with a diameter tape on all juniper and pinon trees. Mean DBH, DRC, and density data were used to calculate basal area (BA) on each plot according to the formula S((DBH/2)2 x 3.14) x 10. 75 N Legend =1 Mixed Conifer Ponderosa Pine/ Gambel Oak Ponderosa Pine/ Pinyon- Juniper Pinyon/ Juniper Aspen Forest Meadows i Water Scale in Miles 0.5 0 1 Figure 2.— Vegetative cover types on the Chevelon Study Area, based upon Terrestrial Ecosystem Surveys (Laing et al. 1989). Canopy coverage of forbs, grasses, shrubs, de- ciduous trees, conifer trees, and rocks was esti- mated along 4 25-ft line intercept transects (Canfield 1941). The first transect was oriented randomly, radiating from site center. The 3 remain- ing transects were each oriented 90° from the pre- ceding transect. We estimated canopy coverage in 3 height categories: 1) 0-17.9, 2) 18-35.9, and 3) 36-72 in. We estimated overhead canopy density with a spherical densiometer (Strickler 1959) at 4 points, 37.2 ft from the feeding site center, along the same bearing as the line intercept transects. We averaged the 4 values to calculate a mean canopy density for each site. We ocularly estimated the distance to the near- est canopy opening from each site center. We de- fined canopy opening as any horizontal gap in the overstory canopy that was greater than 100 ft . We also ocularly estimated the dimensions and calcu- lated the area of the canopy opening. 76 We recorded measurements on the same habi- tat parameters at 103 random plots to compare with feeding sites. Computer generated Universal Transverse Mercator coordinates were plotted on 7.5' U. S. Geological Survey maps. We located each of these points on the ground, and then paced a random distance (<300 ft) on a random bearing to facilitate random plot center placement. This pro- cedure was used to avoid any biases associated with initial random point location. Random plots were measured during the same season as feeding sites. Use of pinon-juniper cover type was analyzed using Chi-square contingency table analysis, Bon- ferroni confidence intervals (Neu et al. 1974, Byers et al. 1984), and Jacobs' D selection index (Jacobs 1974). Because most data violated normality as- sumptions, we used the Mann-Whitney U statistic (Zar 1984) to test for differences between feeding sites and random plots. Differences were consid- ered significant if P < 0.10. Further, differences were not considered significant if both group me- dians were 0, even though P < 0.10. RESULTS Seventy Merriam's turkey were captured and radio-instrumented for winter habitat study (Table 1). Twenty- two feeding sites and 17 random plots were located within pinon-juniper cover type. These comprised 9% of the turkey winter feeding sites and 16.5% of the random plots located on the CSA. Thus, turkeys used the pinon-juniper cover type less than its availability within the CSA (X2 = 4.017, 1 df, P = 0.044, Jacobs' D = -0.285). Table 1.— Age and sex of Merriam's turkey monitored for winter Male Female Year Juvenile Adult Juvenile Adult Total 1988 2 0 4 3 9 1989 0 0 0 1 1 1990 0 1 4 18 23 1991 1 9 3 14 27 1992 6 2 0 2 10 Total 9 12 11 38 70 Table 2. — Median values for feeding sites and random plots in pinon-juniper habitats on the CSA and Mann-Whitney U P values. Habitat Component8 Feeding Site Random Plot P Gambel Oak Tree Density 35.0 0.0 0.056 Gambel Oak Seedling Density 170.0 30.0 0.074 Canopy Density 28.4 40.0 0.043 Distance to Opening 8.0 0.0 <0.001 Opening Size 1375.0 3300.0 0.020 Forb Canopy Cover 0.0 0.1 0.051 Tree densities are reported on a per acre basis, canopy density and cover as a percent, distance in ft, and size in ftz. Feeding sites had greater densities of Gambel oak seedlings and trees (P = 0.074 and 0.056, re- spectively) than at random plots (Table 2). Canopy density was less (P = 0.043) at feeding sites than at random plots (Table 2). Feeding sites were further (P < 0.001) from canopy openings as well (Table 2). The size of the canopy openings adjacent to feed- ing sites was also smaller (P = 0.020) than those found at random plots (Table 2). Canopy cover of forbs was lower (P = 0.051) at feeding sites than at random plots (Table 2). No other difference was found between sites based upon canopy cover (Table 3). Mean DBH or DRC did not differ for any tree species between sites, nor did BA or any other tree density (Table 3). Table 3. — Median0 values for pinon-juniper habitats on the CSA Habitat Component0 Median P Grass Canopy Cover 2.6 0.209 0-17.9 in Deciduous Canopy Cover 0.0 0.013 18-35.9 in Deciduous Canopy Cover 0.0 0.041 36-72 in Deciduous Canopy Cover 0.0 0.017 0-17.9 in Conifer Canopy Cover 2.6 0.260 18-35.9 in Conifer Canopy Cover 3.6 0.812 36-72 in Conifer Canopy Cover 9.2 0.874 0-17.9 in Shrub Canopy Cover 0.3 0.561 18-35.9 in Shrub Canopy Cover 0.1 0.893 36-72 in Shrub Canopy Cover 0.0 0.199 Downed Wood Canopy Cover 2.1 0.657 Rock Canopy Cover 0.4 0.379 Mean Pinus ponderosa DBH 8.5 0.225 Mean Juniperus DRC 10.6 0.821 Mean Pinus edulis DRC 6.4 0.124 Mean Quercus gambelii DBH 3.4 0.539 BA 84.4 0.610 Pinus ponderosa Tree Density 20.0 0.291 Pinus ponderosa Seedling Density 10.0 0.917 Juniperus Tree Density 70.0 0.910 Juniperus Seedling Density 160.0 0.788 Pinus edulis Tree Density 75.0 0.394 Pinus edulis Seedling Density 180.0 0.125 aMedians are presented across groups because no differences between groups were detected. bTree densities are reported on a per acre bases, canopy cover as a percent, DBH and DRC in inches, and BA as ft2 lac. DISCUSSION Although, the pinon-juniper cover type did not comprise a large proportion of turkey winter range, we believe that this cover type may be essential to turkeys during severe winter conditions that in- clude deep (>1 ft) snow or poor food availability. Our study suggests that the pinon-juniper cover type may be used less than available during some winters. We speculate that this occurred during our study because winter weather conditions were mild and winter food availability adequate. Cer- tainly, not all winters are mild nor mast crops plen- 77 tiful. In fact, Shaw (1986) implicated severe winters as a potential cause for a statewide decline in Ari- zona's turkey populations following the winter of 1978-79. During such winters, the suitability of the pinon-juniper cover type is of paramount impor- tance because turkeys rely on it for winter survival (Ligon 1946, Hoffman et al. 1993). This cover type is apparently used for emergency winter feeding during periods of low food availability or deep snow accumulations in adjacent habitats (Reeves 1954, Hoffman et al. 1993). Consequently, the pi- non-juniper cover type can be critical to overwinter survival of Merriam's turkey during severe winters, even though it may receive limited use during mild winters. Habitats <1 mi from roost sites receive the greatest proportion of winter use by turkeys (Wakeling and Rogers In press). Consequently, those areas < 1 mi from potential roost sites, such as the ponderosa pine cover type, a pine stringer habitat, or other suitable cover type (e.g. cotton- wood [Populus fremontii] riparian corridors), are probably the most critical portions of the pinon- juniper cover type to turkeys during severe win- ters. If mamtaining or increasing Merriam's turkey population size is a management goal, these por- tions of the pinon-juniper cover type should be managed to provide suitable feeding habitat and emergency winter range during all years. Using this strategy, adequate reserves will be available during years with adverse weather or poor food supplies. In our study, higher densities of Gambel oak trees were found in feeding habitat selected by turkeys within the pinon-juniper cover type. Be- cause winter diets tend to be comprised mostly of mast (Reeves and Swank 1955, Laudenslager and Flake 1987, Rumble 1990), mast producing species may be used to identify potential feeding habitat. Selection of feeding habitat characteristics may be explained either by plant phenological devel- opment or turkey behavior. Habitats that contained increased densities of deciduous trees would, by their nature, have decreased canopy density dur- ing winter before spring budding and leaf devel- opment. Because feeding sites were selected in higher densities of Gamble oak, they had lower canopy densities than random plots. Feeding sites were selected under trees, where mast crops fre- quently collect. Because feeding sites were selected under trees, feeding turkeys would not be located in openings while feeding. This behavior would explain the greater distance that feeding sites were selected from openings than that distance between random plots and openings. Turkeys generally select smaller openings for feeding activities in southwestern habitats during most seasons (Mollohan and Patton 1991, Hoffman et al. 1993). Smaller openings may be selected because turkeys can reach cover faster if they detect a perceived predator. Our research has identified 2 management strategies that may favor Merriam's turkey using pinon-juniper habitats. First, because winter food supplies are important to turkeys (Wakeling and Rogers In press), protecting and increasing mature mast producing species, such as Gambel oak, juni- per, and pinon trees, at >90 BA (all trees within stand) <1 mi from known or potential roost trees would be beneficial. Mature oaks, with regenera- tion in the understory, appear to be favored by turkeys. Additionally, mature alligator junipers were frequently found within the feeding site. Both Gambel oak and alligator juniper mast play an important role in turkey winter diets in the area (Wakeling and Rogers In press). Second, fuel wood harvests and prescribed burns that protect and enhance openings <0.06 ac would favor turkey populations by retaining more suitable winter feeding habitat. We speculate that the density of these small openings should not exceed 2/ac, thus maintaining about 10% of the area in openings as recommended by Hoffman et al. (1993). LITERATURE CITED Bailey, W, Dennett, D., Gore, H., Pack, J., Simpson, R., and Wright, G. 1980. Basic considerations and general recommendations for trapping the wild turkey. Pro- ceedings of the National Wild Turkey Symposium 4:10-23. Byers, C. R., Steinhorst, R. K., and Krausman, P R. 1984. Clarification of a technique for analysis of utilization- availability data. Journal of Wildlife Management 48:1050-1053. Canfield, R. H. 1941. Application of the line interception method in sampling range vegetation. Journal of For- estry 39:388-394. Glazener, W. C, Jackson, A. S., and Cox, M. L. 1964. The Texas drop-net turkey trap. Journal of Wildlife Man- agement 28:280-287. Hoffman, R. W, Shaw, H. G., Rumble, M. A., Wakeling, B. E, Mollohan, C. M., Schemnitz, S. D., Engel-Wilson, R., and Hengel, D. A. 1993. Management guidelines for Merriam's wild turkeys. Colorado of Division Wildlife, Division Report Number 18. 24. Jacobs, J. 1974. Quantitative measurement of food selec- tion. Oecologia 14:413-417. Laing, L., Ambos, N., Subirge, T, McDonald, C, Nelson, C, and Robbie, W. 1989. Terrestrial ecosystem survey 78 of the Apache Sitgreaves National Forests. USDA Forest Service. 453. Larson, M., and Moir, W. H. 1986. Forest and woodland habitat types (plant associations) of southern New Mexico and central Arizona (north of the Mogollon Rim). USDA Forest Service, Region 3, Albuquerque. 131. Laudenslager, S. L., and Flake, L. D. 1987. Fall food habits of wild turkeys in south-central South Dakota. Prairie Naturalist 19:37-40. Ligon, J. S. 1946. History and management of Merriam's wild turkey. New Mexico Game and Fish Depart- ment, Santa Fe. 84. Mollohan, C, and Patton, D. R. 1991. Development of a habitat suitability model for Merriam's turkey. Ari- zona Game and Fish Department Technical Report Number 9. 217. National Oceanic and Atmospheric Administration. 1991. Arizona climatological data. Volume 95. Neu, C. W., Byers, C. R., and Peek, J. M. 1974. A tech- nique for analysis of utilization-availability data. Journal of Wildlife Management 38:541-545. Phillips, F. 1982. Wild turkey investigations for Bill Wil- liams Mountain area. Special Report Number 13, Ari- zona Game and Fish Department, Phoenix. 50. Reeves, R. H. 1953. Habitat and climatic factors as an influence on Merriam's turkey. Arizona Game and Fish Department, Phoenix. 10. Reeves, R. H. 1954. Merriam's turkey management research. P-R Quarterly 14:7-9. Reeves, R. H., and Swank, W. G. 1955. Food habits of Merriam's turkeys. Arizona Game and Fish Depart- ment, Phoenix. 17. Rumble, M. A. 1990. Ecology of Merriam's turkeys (Meleagris gallopavo merriami) in the Black Hills, South Dakota. Ph. D. Dissertation, University of Wyoming, Laramie. 169. Schorger, A. W. 1966. The wild turkey: its history and domestication. University of Oklahoma Press, Nor- man. 625. Scott, V. E., and Boeker, E.L. 1977. Responses of Mer- riam's turkey to pinyon-juniper control. Journal of Range Management 30:220-223. Shaw, H. G. 1986. Impact of timber harvest on Merriam's turkey populations. A problem analysis report. Ari- zona Game and Fish Department, Phoenix. 18. Shaw, H. G., and Mollohan, C. M. 1992. Merriam's tur- key. Pages 331-349 in Dickson, J. G. The wild turkey: biology and management. Stackpole Books, Harris- burg, PA. Strickler, G. S. 1959. Use of the densiometer to estimate density of forest canopy on permanent sample plots. Research Note PNW-180, Pacific Northwest Forest and Range Experiment Station. 5. Wakeling, B. F. 1991. Population and nesting characteris- tics of Merriam's turkey along the Mogollon Rim, Arizona. Arizona Game and Fish Department, Tech- nical Report Number 7. 48. Wakeling, B. F, and Rogers, T D. In press. Winter habitat relationships of Merriam's turkey along the Mogollon Rim, Arizona. Arizona Game and Fish Department, Technical Report. Zar, J. H. 1984. Biostatistical analysis. Prentice Hall, Englewood Cliffs, New Jersey. 718. 79 Wildlife Associations in Rocky Mountain Juniper in the Northern Great Plains, South Dakota Mark A. Rumble1 and John E. Gobeille2 ABSTRACT. — Rocky Mountain juniper is an important habitat component in the northern Great Plains. These woodlands provide vertical and horizontal vegetative structure that enhances wildlife use. Ecological approaches to managing habitats require understanding relationships between wildlife species and succession in plant communities. We determined bird, small mammals and large mammals habitat use in serai stages of Rocky Mountain juniper woodlands along the Missouri River in South Dakota. Fifty-three bird species occurred in these woodlands. Bird species diversity averaged 2.7 across 24 study sites and we tallied an average of 4.6 bird species during a 3-day sample session at each site. Black-billed Magpies and Blue Jays were the only tree-nesting species whose abundance differed statistically among serai stages of juniper. Trends in the data suggested tree and shrub nesting guilds, total bird abundance, bird species diversity, and birds species richness increased in early and late serai stages. Snags and cavity-nesting species were rare in all serai stages. Northern Flickers were more abundant in late serai juniper; House Wrens were more abundant in intermediate serai juniper. Ground- nesting species declined from low serai to high serai stages. White-footed mice, deer mice, prairie voles, total small mammal abundance, and small mammal species richness were higher in the intermediate serai stage of juniper. Eastern cottontail abundance was greatest in late serai juniper. Trends in deer use suggested higher use of early and late serai stages of juniper woodlands. Patch size, juxtaposition of other woodlands, and animal home range size likely influenced wildlife abundance in serai stages of Rocky Mountain juniper. INTRODUCTION Rocky Mountain juniper (Juniperus scopulorum) woodlands exist in scattered stands in the Missouri River basin. This species is near the eastern limit of its range (Noble 1990). Juniper woodlands occur in highly eroded and rugged terrain. There is evi- dence of past harvest of larger trees for fence posts (Hansen et al. 1984). Larger trees occur in areas protected from fire or inaccessible to firewood cutting. Rocky Mountain juniper does not sprout following fire or cutting (Noble 1990) and the areal 1 Research Wildlife Biologist, Rocky Mountain Forest and Range Experiment Station, Rapid City, SD. Headquarters is in Fort Collins, CO. Wildlife Biologist, Confederated Salish and Kootenai Tribe, Flat- head Reservation, Pablo, MT. extent of juniper woodlands in the northern Great Plains has increased with suppression of fires (Wright etal. 1979). Grasses and forbs of the mixed -grass prairie are common in juniper woodlands. Little-seed rice- grass (Oryzopsis micrantha) is largely restricted to juniper woodlands in this region and it is often dominant in the understory (Hansen et al. 1984, Girard et al. 1989). Vegetative descriptions of "climax" stands of Rocky Mountain juniper in the northern Great Plains occur in Hansen et al. (1984) and Girard et al. (1989). 80 Juniper woodlands are important habitat for wildlife because adjacent grassland habitats lack vegetation structure (Sieg 1991a). Greater vertical structure increases species richness and density of passerine birds (Willson 1974, Roth 1976, Rotenberry and Weins 1980). Juniper woodlands provide habitat necessary for existence of many bird species on the northern Great Plains (Sieg 1991a,b). Juniper woodlands are also important for mammals. Some species of small mammals occur in greater abundance in juniper woodlands in the northern Great Plains (MacCracken et al. 1985a,b; Sieg 1988). Juniper woodlands are important habitats for mule deer (Severson and Carter 1978). The relationships between wildlife and succes- sion in Rocky Mountain juniper have not been examined for this region. The objective of our research was to use a coarse filter approach and examine wildlife abundance in relation to serai stages of Rocky Mountain juniper along the Mis- souri River in South Dakota. STUDY AREA AND METHODS We sampled 24 sites in 3 serai stages of Rocky Mountain juniper (N = 13, 5, and 6 in early, inter- mediate, and late serai stages, respectively) along the Missouri River from 1990 to 1992. A quantita- tive serai stage classification (Uresk 1990) was applied to Rocky Mountain Juniper woodlands to estimate serai stages (unpubl. data, Rocky Moun- tain Experiment Station, Rapid City, SD). The areal extent of juniper stands surrounding sample sites varied. The minimum size was large enough to encompass a 40 X 50 m plot for vegetation samples. Of the sites we sampled, early serai juniper woodlands averaged 326 juniper seedling- saplings/ha (<2.5 cm dbh) and 393 trees/ha (>2.5 cm dbh). Juniper seedling-saplings increased to 2087/ha with >1235 trees/ha in the intermediate serai stage. Seedling-sapling density declined to 526/ha in the late serai stage and tree density aver- aged 1418/ha. Tree density was similar between intermediate and late juniper, but average dbh increased from 7.5 cm to 12.7 cm and tree basal area nearly doubled in late serai juniper. Overstory canopy cover averaged 20% in early serai juniper, 50% in intermediate, and 54% in late serai juniper. Grass cover averaged 42%, 18%, and 34% in early, intermediate, and late serai stages, respectively Little-seed ricegrass averaged 19% cover in late serai juniper. Shrubs averaged 13%, 30%, and 4% cover in early to late serai juniper, respectively. Birds We estimated bird abundance at each site using the variable circular-plot technique (Reynolds et al. 1980). Bird counts were conducted between sun- rise and 1100 hours for three consecutive days during each sample session. In 1990, we completed one sample session in late June. During 1991-92, we completed three sample sessions between May 15 and July 4. During a 5-minute sampling inter- val, we recorded the identity of all birds seen or heard, sex, estimated distance from the sample point, and presence in or out of the woodland. The sampling order of sites was changed daily to en- sure that bird counts occurred throughout the morning. We did not count birds when wind speeds exceeded 10 km/hr or during rainy weather. Mammals We estimated small mammal abundance by trapping small mammals in Sherman live-traps. Two lines of 10 traps spaced at 5-m intervals were established 10 m apart. Traps were baited with commercial bird seed mix and a mixture of peanut butter and rolled oats. We trapped small mammals four consecutive nights during the last week of July or first three weeks of August each year. Trapping periods of 3-5 days are adequate for estimating relative abundance of rodents (Johnston and Keller 1983). Each animal captured was toe-clipped for identification. We estimated medium and large mammals use of juniper woodlands using scent pit surveys (Linart and Knowlton 1975, Conner et al. 1983, Drew et al. 1988). Vegetation and litter were cleared from four 1-m plots and the plot was covered with a fine layer of sifted soil. Two plots were located 25 m above and below the center of the sample site and two, 10 m on either side. A plaster disc soaked in fatty-acid scent (USDA, Animal Pesticide Health Inspection Service, Pocatello, ID) was placed in the center of each plot. We prepared plots during the day and identified tracks (Murie 1974) of animals in the sifted soil the next morning. Each plot was pre- pared again and tracks were recorded on the second day. Scent pellets were removed after each sample session. Three, two-day sample sessions were completed in June, September, and November 1990; four sample sessions (April, June, September, and November) were com- pleted in 1991 and 1992. 81 Analyses RESULTS Detection and density estimates of most birds decline beyond 30 m (Emlen 1971, Verner and Ritter 1988). Only birds within 30 m of the count point and within the juniper woodland were included in our analyses. Although data are reported as density estimates within 30 m of the census point, they should not be construed as absolute estimates of density. It is impossible to assure that all birds within 30 m of the census point were counted (Verner 1985). We summarized data to estimate abundance of birds and guilds (DeGraaf et al. 1991) as averages per site in a hierarchical manner. Data were aggregated to calculate averages across days within sample ses- sion, sample sessions within years, and years within sites. Some bird species occurred on fewer than five sites; these data were not presented. However, we included data from all birds within 30 m in calcula- tions of guild abundance, species richness, and spe- cies diversity. Bird species diversity was calculated using the Shannon-Wiener formula from average abundance per site. Bird species richness is the num- ber of species at each site per sample session. Small mammal abundance is estimated as the av- erage number of unique individuals per night at each site. Small mammal species richness is the number of species captured at each site. Medium and large mammals are the average number of visits each night for each site. Our data did not meet homogeneity of variance and normality assumptions of parametric statistics. We used a multi-response permutation procedure (MRPP) (Mielke 1984) to test hypotheses of no differ- ences of bird abundance, guild abundance, species diversity, and species richness among serai stages. MRPP was also used to make comparisons of small and large mammal abundance among serai stages. We used a a < 0.20 for tests of differences among serai stages. Our objectives were to provide manag- ers with information regarding wildlife relationships to serai stages in juniper woodlands. Sample sites were selected along 320 km of the Missouri River without regard to size of woodlands or landscapes. Woodland area and juxtaposition of other vegetation types add variability to wildlife associations with serai stages. We did not select more stringent levels of statistical significance because Type II errors were as important as Type I errors. Common and scientific names of birds and mammals and their occurrence are in Tables 1 and 2. We observed 53 bird species in juniper wood- lands. Many species were rare or occurred on fewer than 5 sites. Bird species diversity averaged 2.7 for juniper woodlands regardless of serai stage; bird species richness averaged 4.6 per 3-day sample session regardless of serai stage. Seven small mammal species were trapped and 11 other mam- mal species visited scent pits. Small mammal spe- cies richness averaged 2.2 per site. Bird Responses To Serai Stages Black-billed Magpies were more abundant (P = 0.13) in intermediate than early serai juniper (fig. 1). Despite a higher average, abundance of magpies in late serai juniper did not differ from other serai stages. Blue Jays were more (P = 0.03) abundant in late serai juniper than the early or intermediate serai stages. Tree-nesting species tended to be more common in early and late serai stages than the intermediate serai stage. Northern Flickers were more (P < 0.08) abundant in late serai juniper (fig. 2) and intermediate serai juniper had more (P = 0.01) House Wrens than late or early serai stages of juniper. Abundance of cavity- nesting species showed trends toward increasing from early to late serai stages in juniper. No differ- ences were apparent among birds in the shrub nesting guild (fig. 3). Common Yellowthroats and Rufous-sided Towhees showed trends toward higher abundance in later serai stages (fig. 4). Vesper Sparrows were more (P = 0.03) abundant in early serai juniper, while Field Sparrows, and lark sparrows showed trends toward greater abun- dance in early serai juniper. Brown-headed Cowbirds were marginally more abundant (P = 0.21) in early serai juniper and declined in intermediate and late serai stages (fig. 5). Trends suggested total bird abundance and bird species richness reflected abundance of the tree nesting guild, but not significantly. Bird species diversity did not differ among serai stages (P = 0.90). 82 Table 1. — Common name, genus, species, and nesting guild of birds observed in juniper woodlands along the Missouri River, South Dakota, 1990-1992. Serai Stage2 oominon nams uen us/ species •j una Early I n 18 r m 9u i a I© 1 ata Killdeer Charadrius vociferus la A Y X Turkey Vulture Cathartes aura T X X Prairie Falcon Falco mexicanus i L X Ring-necked Pheasant Phasianus colchicus G X X X Wild I urkey Meleagris gallopavo la v X Mourning Dove Zenaida macroura 1 v A Y A Y X Teiiow-DUiea ouckoo L/Occyzus wTiericcinus C o A DlaCK-Ullieu OUL-KUO L/ULcyzuo GryinrufjutturTius Q O V A onun-earea v^wi no/O 1 luff Iff IGUS va Y A Y A m /~\ n Klin o \a/Lv ooiTiiTion iNigninaWK PhnrWo//oo mirmK Y A Y A Y A Northern FIick©r \sOiapies auraius Y A Y A tasiern r\ingDira Ti/ronni IP h/ranni ip /yr annus lyrannus T 1 Y A Y A Y A Western Kingbird lyrannus vemcaiis T 1 Y A Y A Great Crested Flycatcher Myiarchus crinitus Y X Y A Empidonax Flycatchers Empidonax species T v X Tree Swallow Tachycineta bicolor C v X v X Northern Rough-winged Swallow Stelgidopteryx serripennis i L X X Clrrr bwallow Hirundo pyrrhonota i L v X v X v X blue Jay Cyanocitta cristata T v X v X v X black-billed Magpie Pica pica 1 v X v X v X Black-Capped Chickadee Parus atricapillus Y X Y X Y A VV 1 1 1 Lc-Ul caoltrU INUinalCn o/uH ctuoiinensis Y A House Wren Troglodytes aedon Y X Y X Y A Wood Thrush Hylocichla mustelina o o Y A Y A American Robin Turdus migratorius 1 Y X Loggerhead Shrike Lanius ludovicianus O o V X V X Y A oray oaiDiro Dumetella carolinensis o o Y X orown i nrasner Toxostoma rufum o o Y X Y X Y A Cedar Waxwing Bombycilla cedrorum T X X X Red-eyed Vireo Vireo olivaceus 1 v X Warbling Vireo Vireo gilvus 1 v X Teiiow waruier Dendroica petechia o o v X Y A ijeoiniypis zncnas va Y A Y A Y A Mmencan neasian Setophaga ruticilla T 1 Y A rneuciicus meianocepnaius T 1 Y A Y A Rliia ^rnokool/ Guiraca caerulea 1 Y A Y A inuigo Duruing Da a A A r"f f-^ A Ail / A AAA rasserina cyanea c o Y A Rl iti"M ic.cinan 1 ^VA/h^a nU lUUb-blvJcU 1 OWMcc ripiio eryinropninai mus va Y A Y A Y A orassnopper oparrow Ammodramus savannarum la Y Vesper Sparrow Pooecetes gramineus G v X Y X Y A Savannah Sparrow Passerculus sandwichensis Q v X Y A Song Sparrow Melospiza melodia G X Lark Sparrow Chondestes grammacus A, s X v X Y A Field Sparrow Spizella pusilla G X Y X Y A Chipping Sparrow Spizella passerine G X Y A Y A Lark Bunting Calamospiza melancorys G X Western Meadowlark Sturnella neglecta G X X X Red-winged Blackbird Agelaius phoeniceus S X X X Brown-headed Cowbird Molothrus ater P X X X Common Grackle Quiscalus quiscula T X X X Northern Oriole Icterus galbula T X House sparrow Passer domesticus C X American Goldfinch Carduelis tristis T X X X Gu/7d symbols— T = tree nesting; C = cavity nesting; S = shrub nesting; G = ground nesting; L = ledges, cliffs, and crevices; and P = nest parasite. X indicates present in serai stage. 83 Table 2. — Common and scientific names and occurrence by serai stage of small and large mammal species captured or visiting scent pits in juniper woodlands along the Missouri River, South Dakota, 1990-1992. Serai Stage1 Common name Genus/species Early Intermediate Late Thirteen-lined ground squirrel Spermophilus tridecemlineatus X Prairie vole Microtus ochrogaster X X X Meadow vole Microtus pennsylvanicus X X X House mouse Mus musculatus X X X Hispid pocket mouse Perognathys hispidus X White-footed mouse Peromyscus leucopus X X X Deer mouse Peromyscus maniculatus X X X Coyote Canis latrans X X X Porcupine Erethizon dorsatum X Bobcat Lynx rufus X White-tailed jackrabbit Lepus townsendii X Striped skunk Mephitis mephitis X X X Deer Odocoileus spp. X X X Raccoon Procyon lotor X X X Eastern cottontail Sylvilagus floridanus X X X Badger Taxidea taxus X X Red fox Vulpes vulpes X 1 X indicates present in serai stage. Mourning Dove Eastern Kingbird Black-billed Magpie 3.0 2.5 2.0 1.5 1.0 0.5 0 CD CO 0.6 1.0 0.8 0.6 0.4 0.2 0.49 0.42 0.35 0.28 0.21 0.14 0.07 AB Northern Flicker Black-capped Chickadee ll Blue Jay O CD I CO "D 0.5 0.4 0.3 CD 0.2 CD U) 0.1 CO CD 5 Black-headed Grosbeak Common Grackle 0.24 I 0.20 0.16 0.12 0.08 0.04 I .1. 0 American Goldfinch Tree Nesting Guild Early Late Intermediate 0.5 r 0.4 0.3 0.2 CD CO 01 +-> o ® 0 ■o m cd O) 0.25 CO § 0.20 < 0.15 0.10 0.05 0 0.8 0.6 0.4 0.2 House Wren B Cavity Nesting Guild 0.8 0.6 0.4 0.2 Early Late Intermediate Early Late Intermediate Early Late Intermediate Early Late Intermediate Serai Stage Figure 1. — Average abundance (birds/ha) of tree-nesting species among serai stages of Rocky Mountain juniper along the Mis- souri River, South Dakota 1990-92. Standard errors are indi- cated by vertical lines. Letters above abundance bars indicate significant differences among serai stages a < 0.20, MRPP test. Serai Stage Figure 2. — Average abundance (birds/ha) of cavity-nesting species among serai stages of Rocky Mountain juniper along the Mis- souri River, South Dakota 1990-92. Standard errors are indi- cated by vertical lines. Letters above abundance bars indicate significant differences among serai stages a < 0.20, MRPP test. 84 Rufous-sided Towhee Common Yellowthroat 1.5 1.2 0.9 0.6 £ 0.3 0.30 0.25 0.20 0.15 0.10 0.05 0 - Loggerhead Shrike Vesper Sparrow Yellow Warbler I Early Late Intermediate 1.0 0.8 0.6 0.4 0.2 Brown Thrasher Shrub Nesting Guild Early Late Intermediate Serai Stage Figure 3. — Average abundance (birds/ha) of shrub-nesting species among serai stages of Rocky Mountain juniper along the Mis- souri River, South Dakota 1990-92. Standard errors are indi- cated by vertical lines. Letters above abundance bars indicate significant differences among serai stages a < 0.20, MRPP test. Ground Nesting Guild Early Late Intermediate Early Late Intermediate Early Late Intermediate Serai Stage Figure 4. — Average abundance (birds/ha) of ground-nesting species among serai stages of Rocky Mountain juniper along the Missouri River, South Dakota 1990-92. Standard errors are indicated by vertical lines. Letters above abundance bars in- dicate significant differences among serai stages a < 0.20, MRPP test. Mammal Responses To Serai Stages White-footed mice were more (P < 0.12) abun- dant in intermediate and late serai stages of juniper than the early serai stage (fig 6). Deer mice were the most common small mammal and abundance in intermediate serai juniper was greater (P < 0.09) than early or late serai juniper. Meadow voles were more (P < 0.06) abundant in early than in- termediate or late serai stages. Prairie voles were more (P = 0.01) abundant in intermediate than early serai juniper; abundance in late serai juniper did not differ from the previous. Intermediate serai juniper woodlands had more (P < 0.03) small mammals -and greater (P < 0.11) species richness than early or late serai stages. Hispid pocket mice occurred in low abundance in early serai juniper sites, but no differences were apparent among serai stages. Visitations by coyote, raccoon, striped skunk, and badger did not differ among serai stages (fig. 7). Trends suggest the latter three may have used late serai juniper woodlands more than other serai stages. No trends were evident in deer use among serai stages of juniper. Cottontail abundance in- creased (P = 0.04) as succession progressed from early to late serai stages. DISCUSSION We expect juniper woodlands to expand in dis- tribution with a trend toward late serai stages because of fire suppression. Grazing has little 85 Brown-headed Cowbird Total Birds Bird Species Diversity Bird Species Richness Early Late Early Late Intermediate Intermediate Serai Stage Figure 5. — Average abundance (birds/ha) of Brown-headed Cowbirds, total birds, bird species diversity, and bird species richness among serai stages of Rocky Mountain juniper along the Missouri River, South Dakota 1990-92. Standard errors are indicated by vertical lines. Letters above abundance bars in- dicate significant differences among serai stages a < 0.20, MRPP test. direct impact on juniper (Severson and Boldt 1978), but excessive grazing encourages expansion of this woodland (Wright et al. 1979). Rocky Mountain juniper woodlands are "edge" or "island" habitats in the prairie. Wildlife populations were influ- enced by serai stage as well as stand size (Johns 1993), dispersion, and juxtaposition of other woodlands on the landscape. Birds Members of nesting guilds share some resource requirements, but may differ in other critical re- source needs or ability to adjust to alterations of resources (Verner 1984). Managing one or a few indicator species within a guild may not represent all other guild members (DeGraaf and Chadwick 1984, Block etal. 1987). White-footed Mouse Deer Mouse Meadow Vole Early Late Ear,y Late Intermediate Intermediate Serai Stage Figure 6. — Average unique individuals per night of small mammals among serai stages of Rocky Mountain juniper along the Mis- souri River, South Dakota, 1990-1992. Standard errors are in- dicated by vertical lines. Letters above abundance bars indi- cate significant differences among serai stages a < 0.20, MRPP test. Direct linkage between habitat associations of rare species and occurrence in serai stages cannot be made. However, barring better data, our data provide managers with starting points in their habitat assessments. Juniper woodlands add vegetation density and structure to the northern Great Plains. Bird com- munity patterns are associated with vegetation structural diversity (Roth 1976, Rotenberry and Wiens 1980, Sabo and Holmes 1983). Bird occur- rence and abundance in juniper woodlands is greater than in prairie grasslands (Sieg 1991a) and varies with serai stages. Most vegetation structure was provided by the tree canopy and there were more tree nesting birds than birds in other nesting guilds in juniper woodland. Ground-nesting species that use trees for perches or feeding also benefited from juniper woodlands. 86 Vegetative conditions in intermediate serai ju- niper are too dense for ground nesting birds, and not yet favorable for the tree-nesting species. High numbers of juniper saplings and trees result in a dense, uniform stand of small trees in the inter- mediate serai stage. Late serai juniper stands have fewer saplings, taller trees, greater canopy volume, and horizontal patchiness. This structural diversity in vegetation increases abundance and richness (Willson 1974, Holmes et al. 1979, Anderson et al. 1983). Cavity-nesting species were uncommon in ju- niper woodlands because cavities and snags were rare. Northern Flickers was the only woodpeckers in any serai stage of juniper. Black-capped Chicka- dees were the most abundant cavity-nesting spe- cies in our study and another in western South Dakota (Sieg 1991a). No snags >38 cm dbh oc- curred in the sites we sampled and snags 13 - 38 cm Coyote Striped Skunk 0.5 0.4 0.3 0.2 0.1 0 •4— < _c O) Z ~Q 0.30 W 0.25 £ 0.20 O 0.15 B o.io 0.4 0.3 0.2 - 0.1 mm Raccoon Badger CO 0.05 CD 0 Ui CD CD > < 1.0 0.8 0.6 0.4 0.2 ■I 0.10 0.08 0.06 (■ 0.04 0.02 0 I Deer Eastern Cottontail 1.5 1.2 0.9 0.6 0.3 Early Late Intermediate Early Late Intermediate Serai Stage Figure 7. — Average visitations per session to scent pits by inter- mediate and large mammals among serai stages of Rocky Mountain juniper along the Missouri River in South Dakota, 1990-1992. Standard errors are indicated by vertical lines. Letters above abundance bars indicate significant differences among serai stages a < 0.20, MRPP test. dbh occurred only in early (0.4/ha) and late serai juniper (0.8/ha). Suitable snags for cavity nesters were usually species other than juniper. Cavity nesting birds were uncommon in southwest pihon- juniper woodlands if pine trees (P. mono-phyla and P. edulis) were absent (Balda and Masters 1980). Other cavity-nesting species, such as House Wrens, probably nested in nearby green ash or cotton- wood trees. The Loggerhead Shrike is being considered for listing under the Endangered Species Act and is considered threatened or endangered in several states (Finch 1992, Smith and Kruse 1992). Logger- head Shrikes prefer open habitats with scattered perch sites (Finch 1992) typical of the early serai stages of juniper on the northern Great Plains. Vesper Sparrows, Field Sparrows, and Lark Sparrows are ground -nesting birds associated with scattered shrub or woodlands. Abundance of these species declined (some not significantly) from early to late serai stages. Reduced ground cover for nesting from early to late serai stages may have been a factor affecting use of juniper wopdlands by these species. Common Yellowthroats and Rufous- sided Towhees are associated with tall shrubs or small trees that may account for the insignificant increase in late serai juniper. Many of the birds in the ground nesting guild using juniper woodlands are prairie species and do not require woodlands. Brown-headed Cowbirds have been implicated in the decline of song birds in the United States and are obligate nest parasites (Brittingham and Temple 1983). Cowbird abundance was correlated with abundance of both tree (r = 0.6, P < 0.01) and ground (r = 0.4, P = 0.03) nesting birds in this study. Cowbirds prefer open scattered woodlands (DeGraaf et al. 1991) and may have found nests easier in the open canopy early serai stages of juniper. Mammals Species richness of small mammals in Rocky Mountain juniper was low, but low species rich- ness may not be uncommon in northern Great Plains. Other reports show only 6-9 species of small mammals (MacCracken et al. 1985b, Hodorff et al. 1988, Sieg 1988). Abundance and species richness of small mammals contrasted data for birds; the intermediate serai stage supported more small mammals. Reduced variability and stronger associations of small mammal to serai stages of juniper than birds resulted from their resident status and small home ranges. 87 Deer mice are widespread generalist rodents in North America (Baker 1968). In southeastern Mon- tana woodlands, deer mice were positively corre- lated with understory vegetative cover (MacCracken et al. 1985b). Intermediate serai juniper woodlands in our study had more shrub cover, less grass cover, and greater patchiness in the understory than early or late serai woodlands. White-footed mice are usually associated with riparian habitats in the Great Plains (Andersen and Jones 1971, Hodorff et al. 1988). Rocky Mountain juniper woodlands are not riparian ecosystems. Similar to deer mice, white-footed mice were more abundant in the intermediate serai stage. Ribble and Samson (1987) suggested these two species partition macrohabitats but use similar microhabi- tats. We found them using similar macrohabitats in juniper woodlands. Meadow voles are associated with high vegetative cover (Huntly and Inouye 1987) and shrublands (Snyder and Best 1988). They usually occur in moist meadows, but inhabit upland areas if sufficient vegetation is present (Jones et al. 1983, Sieg 1988). Early serai juniper, where meadow voles were more abundant, had greater understory and grass cover than other serai stages of juniper. Upland habitats with herbaceous cover appear to meet habitat needs of prairie voles (Moulton et al. 1981). Prairie voles partition habitats with meadow voles where their distribution overlaps to prevent competition (Jones et al. 1983) and in the juniper woodlands we studied, meadow voles and prairie voles exhibited opposite trends in abun- dance among serai stages. Hispid pocket mice are associate with upland habitats, bare ground (Jones et al. 1983), shortgrass prairie (Moulton et al. 1981), or early serai grass- lands (McMillan and Kaufman 1994). They select habitats with loamy soils, for burrows (Jones et al. 1983). Suitable habitat conditions for Hispid pocket mice occurred only in early serai stages of Rocky Mountain juniper. Hispid pocket mice also used woodland habitats on the Konza Prairie, Kansas (McMillan and Kaufman 1994). House mice also were uncommon, but occurred in all serai stages of Rocky Mountain juniper. Cottontail abundance increased with habitat changes from early to late serai conditions in juni- per. Cottontails are usually associated with shrub habitats in the Great Plains (Morgan and Gates 1983, Rumble 1989). Cottontails in Oklahoma preferred eastern red cedar (J. virginiana) habitats (Lochmiller et al. 1991). Cottontails prefer escape cover and a clear view (Morgan and Gates 1983). Late serai juniper with low, dense branches and open understory provided this habitat better than other serai stages. Juniper, however, is not a pre- ferred food of cottontails (Lochmiller et al. 1991) and they probably fed on other vegetation. Big game animals are attracted to forests in the northern Great Plains for thermal and hiding cover and browse (Martinka 1968, Wood et al. 1989). Late serai juniper provided vegetative cover overhead and horizontally. Juniper woodlands of Badlands National Park (probably late serai stages) were used extensively by mule deer (Severson and Carter 1978) and were important for mule deer fawns in late summer (Steigers 1981). Most deer in juniper woodlands were mule deer. Raccoons are uncommon in dry upland for- ests and grasslands (Fritzell 1977, Jones et al. 1983). Badgers do not depend on prairie wood- lands; they are inhabitants of grasslands and forest edges (Jones et al. 1983). Coyotes inhabit a variety of habitats. Coyotes' diets reflect prey abundance (Andelt et al. 1987, MacCracken and Hansen 1987, Reichel 1991) and they hunt where prey species concentrate (Reichel 1991). We presume coyotes used juniper woodlands for hunting because of the increased abundance of cottontails and other prey. CONCLUSIONS Early and late serai stages of Rocky Mountain ju- niper provided habitat for birds in all nesting guilds. House Wrens were the only bird that occurred in greater abundance in the intermediate serai stage of juniper. Intermediate serai juniper is a dense uniform stand of trees and lacks the vertical and horizontal structure for most bird species. Some ground -nesting species in early serai juniper are prairie species and do not depend on these habitats. Most small mam- mals were more abundant in the intermediate serai juniper, but meadow voles and Hispid pocket mice appeared to prefer early serai juniper woodlands. Late serai juniper provided the best habitat for cotton- tails. Several species of intermediate or large mam- mals also used Rocky Mountain juniper, but were not associated with any serai stages. These species have large home ranges that encompassed several wood- land sites or types. Once objectives for wildlife spe- cies are developed, techniques that maintain or alter serai stages of juniper woodlands to meet manage- ment objectives can be done. Ensuring sustained biological diversity in Rocky Mountain juniper woodlands will require all serai stages across the landscape. 88 ACKNOWLEDGMENTS Special thanks to A. Clements, C. Erickson, C. Hydock, K. Landers, K. Meskill, M. Meyer, C. Oswald, T. Pella, M. Pennock, G. Proudfoot, R. Young, and G. Soehn for assistance collecting field data. T. Hoag, D. Strickland, and D. Uresk developed juniper serai stage classification. R. King and J. Janssen provided statistical advice. U.S. Army Corps of Engineers provided funding under MIPR 9861-89. South Dakota Game, Fish, and Parks provided logistical and technical support. Tribal Council;, and Chairs of the Cheyenne River Sioux, Lower Brule, and Crow Creek Reservations allowed sampling on tribal lands and several private land owners generously allowed us to conduct research on their land. LITERATURE CITED Andelt, William E, Kie, John G., Knowlton, Frederick F.., Cardwell, Kean. 1987. Variation in coyote diets asso- ciated with seasonal and successional changes in vegetation. Journal of Wildlife Management. 51:273- 277. Andersen, Kenneth W, Jones, J. Knox, Jr. 1971. Mam- mals of northwestern South Dakota. University of Kansas Publications, Museum of Natural History. 19:361-393. Anderson, Bertin W, Ohmart R.D., Rice, J. 1983. Avian and vegetation community structure and their sea- sonal relationships in the Lower Colorado River Val- ley. Condor. 85: 392-405. Baker, Rollin H. 1968. Habitats and distribution. In: John A. King, ed. Biology of Peromyscus (Rodentia). American Society of Mammologists. 2: 98-126. Balda, Russel P, Masters, Nancy. 1980. Avian communi- ties in pinyon-juniper woodland: a descriptive analysis. In: R.M. Degraff, ed. Management of west- ern forests and grasslands for nongame birds. USDA Forest Service, General Technical Report INT-86. In- termountain Forest and Range Experiment Station, Odgen,UT. 146-169. Block, William M., Brennan, Leonard A., Gutierrez, R.J. 1987. Evaluation of guild-indicator species for use in resource management. Environmental Management. 11: 265-269. Brittingham, M.C, Temple, Stanley A. 1983. Have cowbirds caused forest songbirds to decline. Bio- science. 33: 31-35. Conner, Mark C, Labisky, Ronald F.., Progulske, Donald R., Jr. 1983. Scent-station indices as measures of population abundance for bobcats, raccoons, gray foxes, and opossums. Wildlife Society Bulletin. 11: 146-152. DeGraaf, Richard M., Chadwick, N.L. 1984. Habitat classification: a comparison using avian species and guilds. Environmental Management. 8: 511-518. DeGraff, Richard M., Scott, Virgil E., Hamre, R. H. [et al.]. 1991. Forest and range land birds of the United States: natural history and habitat use. USDA, Forest Service, Agriculture Handbook 688. 625 p. Drew, Gary S., Fagre, Daniel B., Martin, David J. 1988. Scent-station surveys for cottontail rabbit popula- tions. Wildlife Society Bulletin. 16: 396-398. Emlen, John T. 1971. Population densities of birds de- rived from transect counts. Auk. 88: 323-342. Finch, Deborah M. 1992. Threatened, endangered and vulnerable species of terrestrial vertebrates in the Rocky Mountain Region. USDA Forest Service, Gen- eral Technical Report RM-215. Rocky Mountain For- est and Range Experiment Station, Fort Collins, CO. 38 p. Fritzell, E.K. 1977. Aspects of raccoon (Procyon lotor) social organization. Canadian Journal Zoology. 56: 260-271. Girard, Michele M., Goetz, Harold, Bjugstad, Ardell, J. 1989. Native woodland types of southwestern North Dakota. USDA, Forest Service Research Paper RM- 281. Fort Collins, CO. 36 p. Hansen, Paul L., Hoffman, George R., Bjugstad, Ardell J. 1984. The vegetation of the Theodore Roosevelt Na- tional Park, North Dakota: a habitat classification. USDA Forest Service General Technical Report RM- 113. Rocky Mountain Forest and Range Experiment Station, Fort Collins, CO. 35 p. Hodorff, Robert A., Sieg, C. Hull, Linder, Raymond L. 1988. Wildlife response to stand structure of decidu- ous woodlands. Journal of Wildlife Management. 52: 667-673. Holmes, R.T., Bonney, R.E., Jr., Pacala, S.W. 1979. Guild structure of the Hubbard Brook bird community: a multivariate approach. Ecology. 60: 512-520. Huntly, Nancy, Inouye, Richard S. 1987. Small mammal populations of an old-field chronosequence: succes- sional patterns and associations with vegetation. Journal of Mammology. 68: 739-745. Johns, Brian W. 1993. The influence of grove size on bird species richness in aspen parklands. Wilson Bulletin. 105: 256-264. Johnston, William C, Keller, Barry L. 1983. An examina- tion of snap-trapping techniques for estimating ro- dent density in high desert. Northwest Science. 57: 194-204. Jones, Knox, Jr., Armstrong, David M., Hoffman, Robert S., Jones, Clyde. 1983. Mammals of the northern Great Plains. University of Nebraska Press. Lincoln, NE. 375 p. Linhart, Samuel B., Knowlton, Frederick F. 1975. De- termining the relative abundance of coyotes by scent station lines. Wildlife Society Bulletin. 3:119-124. Lochmiller, R.L., Boggs, J.F., McMurray, ST., [et al.]. 1991. Response of cottontail rabbit populations to herbicide and fire applications on cross timbers rangeland. Journal of Range Management. 44: 150-155. 89 MacCracken, James G., Uresk, Daniel W, Hansen, Richard M. 1985a. Habitat used by shrews in southeastern Montana. Pairie Naturalist. 59: 24-27. MacCracken, James G., Uresk, Daniel W, Hansen, Richard M. 1985b. Rodent-vegetation relationships in southeastern Montana. Northest Science 59: 272-278. MacCracken, James G., Hansen, Richard M. 1987. Coyote feeding strategies in southeastern Idaho: op- timal foraging by and opportunistic predator? Jour- nal of Wildlife Management. 51: 278-285. Martinka, C.J. 1968. Habitat relationships of white- tailed and mule deer in northern Montana. Journal of Wildlife Management. 32: 558-565. McMillan, Brock R., Kaufman, Donald W 1994. Differ- ences in use of interspersed woodland and grassland by small mammals in northeastern Kansas. Prairie Naturalist. 26: 107-116. Mielke, Paul W. 1984. Meterological applications of permutation techniques based on distance functions. Handbook of Statistics. 4: 813-830. Morgan, Kevin A., Gates, J. Edward. 1983. Use of forest edge and strip vegetation by eastern cottontails. Journal of Wildlife Management. 47: 259-264. Moulton, Michael P, Choate, Jerry R., Bissell, Steven J. [et al.]. 1981. Associations of small mammals on the cen- tral high plains of Colorado. Southwestern Naturalist. 26: 53-57. Murie, Olaus J. 1974. A field guide to animal tracks. Houghton Mifflin Company. Boston, MA. 375 p. Noble, Daniel L. 1990. Rocky Mountain juniper. In: Russell M. Burns and Barbera H. Honkala, tech. co- ords. Silvics of North America: conifers. Agricultural Handbood Number 654. USDA Forest Service. Washington, D.C. 1:116-126. Reichel, James D. 1991. Relationships among coyote food habits, prey populations and habitat use. Northwest Science. 65: 133-137. Reynolds, R.T., Scott, J.M., Nussbaum, R.A. 1980. A variable circular-plot method for estimating bird numbers. Condor, 82: 309-313. Ribble, David O., Samson, Fred B. 1987. Microhabitat associations of small mammals in southeastern Colo- rado, with special emphasis on Peromyscus (Rodentia). Southwest Naturalist. 32: 291-303. Rotenberry, John T, Wiens, John A. 1980. Habitat structure, patchiness, and avian communities in North American steppe vegetation: A multivariate analysis. Ecology. 61: 1128-1250. Roth, Roland R. 1976. Spatial heterogeneity and bird species diversity. Ecology. 57: 773-782. Rumble, Mark A. 1989. Wildlife associated with scoria outcrops: implications for reclamation of surface- mined lands. USDA Forest Service, Research Paper RM-285. Rocky Mountain Forest and Range Experi- ment Station, Fort Collins, CO. 7 p. Sabo, Stephen R., Holmes, R.T. 1983. Foraging niches and the structure of forest bird communities in con- tasting montane habitats. Condor. 85: 121-138. Severson, Keith E., Boldt, Charles, E. 1978. Cattle, wildlife, and riparian habitats in the western Dakotas. In: Management and use of northern plains rangeland. North Dakota State University, Dickinson. 91-102. Severson, Keith E., Carter, Arthur V. 1978. Movements and habitat use by mule deer in the northern Great Plains, South Dakota. In: Donald N. Hyder, ed. Proceedings of the First International Rangeland Congress. Society for Range Management. Denver, CO. 466-468. Sieg, Carolyn Hull. 1988. The value of Rocky Mountain juniper (Juniperus scopulorum) woodlands in South Dakota as small mammal habitat. In: Szaro, Robert C, Severson, Keith E., Patton, David R., tech coors. Management of amphibians, reptiles, and small mammals in North America. USDA, Forest Service General Technical Report RM-166. Rocky Mountain Forest and Range Experiment Station, Fort Collins, CO. 328-332. Sieg, Carolyn Hull. 1991a. Rocky Mountain juniper woodlands: year-round avian habitat. USDA, Forest Service Research Paper RM-296. Rocky Mountain Forest and Range Experiment Station, Fort Collins, CO. 7 p. Sieg, Carolyn Hull. 1991b. Geographic affinity of bird species associated with Rocky Mountain juniper woodlands and adjacent grasslands in southwestern South Dakota. Pairie Naturalist. 23: 25-33. Smith, Eric L., Kruse, Kipp C. 1992. The relationship between land-use and the distribution and abun- dance of Loggerhead shrikes in south-central Illinois. Journal of Field Ornithology. 63: 420-427. Snyder, Ellen J., Best, Louis B. 1988. Dynamics of habitat use by small mammals in prairie communities. American Midland Naturalist. 119: 128-136. Steigers, William, D. 1981. Habitat use and mortality of mule deer fawns in western South Dakota. Brigham Young University, Provo, UT. 193 p. . Ph.D. thesis. Uresk, Daniel W. 1990. Using multivariate techniques to quantitatively estimate ecological stages in a mixed- grass prairie. Journal of Range Management. 43:282- 285. Verner, Jared. 1984. The guild concept applied to man- agement of bird populations. Environmental Man- agement. 8: 1-14. Verner, Jared. 1985. Assessment of counting techniques. In: R.F. Johnston, ed. Current Ornithology. (2)247-302. Verner, Jared, Ritter, Lyman V. 1988. A comparison of transects and spot mapping in oak-pine woodlands of California. Condor. 90: 401-419. Willson, Mary E. 1974. Avian community organization and habitat stucture. Ecology. 55: 1017-1029. Wood, Alan K, Mackie, R.J., Hamlin, K.L. 1989. Ecology of sympatric populations of mule deer and white-tailed deer in a prairie environment. Montana Department of Fish, Wildlife, and Parks. 97 p. Wright, Henry A., Neuensch wander, Leon F.., Britton, Carlton. 1979. The role and use of fire in sagebrush- grassland and pinyon-juniper plant communities: a state-of-the-art review. USDA Forest Service, General Technical Report INT-58, Intermountain Forest and Range Experiment Station, Odgen, UT. 90 Effects of Fuelwood Harvesting on Small Mammal Populations in a Pinon- Juniper Woodland William H. Kruse1 Abstract. — Small mammal populations have an intricate role in ecosystem function and must be considered a key component of pinon-juniper woodlands. Current management practices not only affect the habitats of small mammals but also the habitat of their specific predators. Trapping small mammals began in 1990, two years prior to woodland harvesting and will continue through the treatment and into the post- treatment years. Data from the pre-treatment period show a 50% increase in the total number of small mammals from 1990 to 1991 across all plots. A similar trend occurred from 1992-1993, demonstrating large variation among years. Little variation in small mammal numbers were found among plots. By the third trapping year (1992), four of the eight units had been cut and by the 1993 trapping season, all eight units were cut. Preliminary results and field observations suggest that harvesting may negatively affect pihon mice populations. Conversely, the harvesting had a more positive effect on the deer mice numbers as well as on species diversity. INTRODUCTION Pinon-juniper woodlands receive a diversity of uses. Previous attempts at managing these woodlands were directed at increasing forage for livestock by removing the overstory. Current management efforts have been redirected to provide winter range, for both wild ungulates and migratory birds, and most recently, to provide fuelwood. As a result, user interest has also been redirected from single- to multiple-use manage- ment. As the multiple-use management concept includes more nonconsumptive natural resources, it begins to resemble an ecosystem management concept. To complement the National Forest's eco- system management approach, the Rocky Moun- tain Forest and Range Experiment Station is cooperating with the Apache-Sitgreaves National Forest in a multi-year research project (Kruse and Perry 1994). The project is located on the Mud Tanks fuelwood management area to examine the effects of fuelwood harvesting on overstory 1 USDA Forest Service, Rocky Mountain Forest end Range Experi- ment Station, Flagstaff, AZ. understory relationships (overstory regeneration as well as forage production), nutrient cycling, soil erosion, runoff, and on selected wildlife populations. This small mammal study is one component of the project. As the public becomes increasingly more concerned about the low levels of management that historically have been provided for pinon- juniper woodlands (Gottfried 1987), broadened and more diverse ecosystem research, in management strategies, is desired. Because these woodlands cover such a large area there is general agreement that they should be managed for multiple uses (Evans 1988). Also, the popularity of converting woodlands, solely for livestock purposes, has declined, partially because of the current interest in ecosystem management and partially because there are fewer easily converted sites available. More importantly however, is the need to maintain natural systems and to understand how management or changes in specific resources alter the function of that natural 91 system (Gottfried 1987). Finally, the recent emphasis on fuelwood harvesting and slash disposal has added a new dimension to managing these woodlands particularly in the context of managing natural systems (Ffolliott et al. 1979). FUELWOOD HARVESTING AND THE SMALL MAMMAL COMMUNITY The increase of fuelwood harvesting has prompted increased concern for assessing the effect of both fuelwood removal and slash disposal, particularly burning, on nutrient cycling, understory production (specifically forages for livestock and other large ungulates), and small mammals in these pihon-juniper woodlands. In Arizona, fuelwood demands increased over 400% between 1973 and 1978 (Ffolliott et al. 1979). Land managers are currently attempting to develop sound silvicultural prescriptions for these woodlands, but basic ecological information needed to support current harvesting plans is often lacking (Gottfried 1987). Fuelwood harvest is the most significant factor affecting the overstory while the least understood management option has been slash deposition (Severson 1986; Baker and Frischknecht 1972). Because previous research has shown that species' populations can be impacted by overstory disturbances (Turkowski and Reynolds 1970), this study will provide quantitative information on small mammals following fuelwood harvest. For example, Severson (1986), in New Mexico, found that total rodent numbers were significantly greater on treated compared to untreated woodlands. Kruse et al. (1979) found that treated areas differed from the woodland in that those rodent species which preferred the woodland condition, decreased in numbers on the treated areas. Whether trends such as these will follow fuelwood harvesting is the focus of this research. Finally, and because small mammals have an intricate role in the food chain as prey for raptors and carnivores, information determined from this study will provide information for those predators as well. STUDY AREA AND METHODS The study is being conducted on the Heber Ranger District, Apache-Sitgreaves National Forest, in central Arizona. Average tree conditions are 23.2 +_ 5.4 m2/ha of basal area, producing 35.3 + 12.7 m3/ha of fuelwood. Sixty- three percent of the trees are pirion. One-seed juniper is the second most common species followed by alligator juniper. Ponderosa pine (Pinus ponderosa) occasionally occurs on moist sites. Average canopy cover is 40% (Laing et al. 1988). Average annual herbaceous and woody plant potential productivity is about 562 kg/ha. The area is relatively flat, dissected by several small ephemeral drainages. Elevations are between 2,000 and 2,060 m. The primary soil subgroups, derived from limestone, are Lithic Ustochrepts, Udic Haplustalfs, and Typic Eutroboralfs. The mean annual precipitation varies between 34 and 46 cm. The overall research project consists of 33 units, 4-ha in size. Thirty 4-ha study units were grouped into 5 blocks each with 6 overstory treatments. Sixteen of these were selected for the small mammal study. These sixteen contained all the overstory conditions represented in the overall research project. The experimental design entails a randomized block layout. Blocks were designated based on similarity of pre-treatment overstory conditions and characteristics. The experimental units entail combinations of burning, no -burning, and cutting, no-cutting. Hence, the small mammal study repeats these four treatments in each of four blocks. Treatments were assigned randomly to each block and were located as conditions permit and therefore were not necessarily contiguous; roads or drainage channels could separate units within a given block. Nevertheless, all treatments for each block are in the same area. Harvesting began in the fall/winter of 1991 and continued up to 24 months. Burning will commence on those units when the slash approaches 2 years old. A 100m X 100m trapping grid was located in the center of each block utilizing about 85% of it. At each grid point, 10m x 10m apart, was placed an 8 X 10 X 25cm Sherman live trap. At every other point, a 10 X 12 X 40cm Sherman live trap was located with the smaller one. The bait was a mixture of chicken scratch and rolled oats. Thus each unit was sampled yearly with 150 traps for 3 nights and 2 days. Demographic and physical measurements were taken and recorded on each animal caught. Each animal was toe clipped and released. Recaptures were noted. Relative abundance and species composition (Table 1) of small mammals was estimated by live- trapping on four overstory treatments: (1) type conversion (where the fuelwood has been harvested, residual trees cut and the slash burned); beer not bun ovei wer con lab: be 92 (2) cut but not burned (where the fuelwood has been harvested but the non-commercial trees are not cut nor is the slash burned); (3) no cut but burned (analogous to a forest fire where the overstory is removed by fire; some fire ladders were cut to facilitate the burning); and (4) the controls (where the units remain untreated) (see Table 2). Small mammal trapping is being conducted once each year (July-August). The null hypotheses, that there are no differences in (1) total number of small mammals or (2) total number of species among treatments, will be tested utilizing years as repeated measures. Table 1. — Small mammal species list from Mud Tanks. Peromyscus maniculatus deer mouse 38% P. truei pinon mouse 48% P. boylii brush mouse 5% Neotoma albigula white throated wood rat 2% N. mexicanus Mexican wood rat 2% Eutamias dorsalis cliff chipmunk 4% Dipodomys ordii Ord's kangaroo rat <1% Microtus pennsylvanicus meadow vole <1% Spermophilus variegatus rock squirrel <1% Table 2. — Treatment schedule for small mammal study units. UNIT BLKS 1990 1991 1992 1993 1994 1995 1996 1997 10-4 III NOTR NOTR NOTR NOTR NOTR NOTR NOTR NOTR 11-4 IV NOTR NOTR NOTR NOTR NOTR NOTR NOTR NOTR 15-4 I NOTR NOTR NOTR NOTR NOTR NOTR NOTR NOTR 20-2 V NOTR NOTR NOTR NOTR NOTR NOTR NOTR NOTR 14-1 III NOTR NOTR NOTR NOTR NCB1 NCB2 NCB3 NCB4 15-3 IV NOTR NOTR NOTR NOTR NCB1 NCB2 NCB3 NCB4 20-4 I NOTR NOTR NOTR NOTR NCB1 NCB2 NCB3 NCB4 21-2 V NOTR NOTR NOTR NOTR NCB1 NCB2 NCB3 NCB4 15-2 III NOTR NOTR NOTR CNB1 CNB2 CNB3 CNB4 CNB5 14-4 IV NOTR NOTR CNB1 CNB2 CNB3 CNB4 CNB5 CNB6 16-4 I NOTR NOTR CNB1 CNB2 CNB3 CNB4 CNB5 CNB6 21-4 V NOTR NOTR NOTR CNB1 CNB2 CNB3 CNB4 CNB5 14-2 III NOTR NOTR CNB1 CNB2 C&B1 C&B2 C&B3 C&B4 15-1 IV NOTR NOTR NOTR CNB1 CNB2 C&B1 C&B2 C&B3 16-3 I NOTR NOTR CNB1 CNB2 C&B1 C&B2 > C&B3 C&B4 21-1 V NOTR NOTR NOTR CNB1 CNB2 C&B1 C&B2 C&B3 NOTR = No Treatment; NCB1 = No Cut, Burned (year); CNB1 = Cut/Not Burned, (year); C&B1 = Cut & Burned, (year) Table 3. — Mean captures for uncut and cut treatments by year. 1990 1991 1992 1993 UC C UC C UC C UC C P. maniculatus 3.8 4.1 18.4 16.6 2.9 7.5 11.8 45.6 P. truei 16.3 12.8 20.1 19.4 12.5 10.4 30.5 20.0 All others 3.0 3.5 6.1 5.5 2.1 3.7 5,6 10.7 Total animals 23.1 20.4 44.6 41.5 17.5 21.6 47.9 76.3 TOTAL CAPTURES Small Mammals 1990-1993 140 i CO £ 120 E 100 ■ | 80 f 60 .i.i CD O 40 1 20 0 1990 1991 1992 1993 Year: Mean: 1990 43.5 +/- 6 1991 86.4 +/- 8 1992 1993 39.1 124.9 +/- 8 +/- 18 Figure 1. — Highly significant annual variation. PRELIMINARY RESULTS Table 3 shows mean captures for the uncut and cut units for the four years of sampling. These data demonstrate the dominance of the two Peromyscus species relative to the total numbers of all captured animals. Populations fluctuated during the first four years of the study. These first analyses show significant differences between years,(p < 0.001). Figure 1 best expresses this yearly variation, as well as the significant interaction between treatment and year factors. Figure 2 demonstrates an analyses on all small mammal captures and shows the similarity between cut and uncut study units prior to harvest. There were no differences among units prior to treatment (Fig. 2). A significant doubling of total population numbers from 1990 to 1991 is evident. Figures 4 and 6 show again the similarity among pre-treatment study units as well as the similarity within each species' population. An increase in the number of captures was similar for both the pinon mouse and deer mouse between 1990 and 1991. More notable, however, was that of the highly significant (p < 0.001) increase in 1991 over 1990 of the deer mouse which contributed the greater portion was contributed of all small mammal captures. 93 Small Mammal Captures before harvest (1990-1991) tt 50 c "§40 f 30 1.20 20 | |||| 10 n ■ I UNCUT CUT (to be) ITOT90CJTOT91 Deer Mouse Captures before harvest (1990-1991) ~ 20 c L. W CD i— CL 03 O C (0 15 10 UNCUT CUT (to be) PEMA90 □ PEMA91 Figure 2. — Annual yearly variation in small mammal captures prior to harvest. Small Mammal Captures before(1991) and after(1993) harvest ~ 80 c leo 100 years), the potential for changes in vegetation cover must be factored into the design. By clarifying the linkages between hydrological and ecological processes, our framework can contribute to the development of a model for predicting these changes in pirion- juniper woodlands. In addition, the framework and the data being collected could prove useful in addressing the overall problem of erosion in pifion-juniper wood- lands. Management strategies for reversing this trend focus on increasing the amount of herba- ceous cover. Our studies of the relationships be- tween vegetation cover and soil moisture could lead to improved methods for doing that. In summary, we believe that the extreme het- erogeneity of semiarid landscapes can be described in terms of variations in the hydrologic properties of the components of these landscapes. The resul- tant variations in vegetation, both in space and over time, have created the need for a model, or framework, that will allow us to predict the rela- tionships between water, sediment, and vegetation dynamics. Several testable hypotheses have al- ready resulted from the framework we are devel- oping. As we test these hypotheses using the data from our network of field studies, we hope to improve our predictive capabilities, and thereby our ability to solve environmental problems in semiarid landscapes. LITERATURE CITED Allen, CD. 1989. Changes in the landscape of the Jemez Mountains, New Mexico. University of California, Berkeley, 346 p. Blackburn, W.H. 1975. Factors influencing infiltration and sediment production of semiarid rangelands in Nevada. Water Resources Research. ll(6):929-937. Bowen, B.M. 1990. Los Alamos Climatology. Los Alamos National Laboratory Report LA-11735-MS. Breshears, D.D.; Johnson, S.R.; Myers, O.B.; Meyer, C.W.; and Martens, S.N.(in progress) Differential use of shallow intercanopy water by Juniperus monosperma and Pinus edulis: coupling of canopy and intercanopy patches in semiarid woodlands. Breshears, D.D. 1993. Spatial partitioning of water use by herbaceous and woody lifeforms in semiarid woodlands. Ph.D. dissertation, Colorado State Uni- versity, Fort Collins, CO. 78 p. Breshears, D.D.; Whicker, F.W.; Hakonson, T.E. 1993. Orchestrating environmental research and assess- ment for remediation. Ecological Applications. 3(4):590-594. Crowe, B.M.; Linn, G. W; Heiken, G; and Bevier, M.L. 1978. Stratigrapy of the Bandelier Tuff in the Pajarito Plateau. Los Alamos Scientific Laboratory Report LA- 7225-MS. Dawson, T.E. 1993. Woodland water balance. Trends in Ecology and Evolution. 8:120-121. Everett, R. L. (ed.) 1986. Proceedings: Pinyon-Juniper Conference. USDA Forest Service, Intermountain Re- search Station, Ogden, Utah, INT-215:581 Gardner, R. H.; Turner, M. G.; Dale, V. H.; O'Neill, R. V. 1992. A percolation model of ecological flows. In: Hansen, A. J.; di Castri, Francesco (eds.), Landscape Boundaries: Concequences for Biotic Diversity and Ecological Flows. Ecological Studies 92, Springer - Verlag, New York. Gosz, J.R. 1992. Gradient analysis of ecological change in time and space: implications for forest management. Ecological Applications. 2(3):248-261. Gosz, J.R.; Sharp, PJ.H. 1989. Broad-scale concepts of climate, topography, and biota at biome transitions. Landscape Ecology. 3(3/4):229-243. Grover, H.D.; Musick, B.H. 1990. Shrubland encroach- ment in southern New Mexico, USA: an analysis of desertification processes in the American Southwest. Climate Change. 17:305-330. Hook, PB. 1992. Individual-plant scale patterns of roots and soil resources in shortgrass steppe. Colorado State University, Fort Collins, CO. Lajtha, K., Getz, J. 1993. Photosynthesis and water-use efficiency in pinyon-juniper communities along an elevation gradient in northern New Mexico. Oecolo- gia. 94:95-101. Lajtha, K.; Barnes, F.J. 1991. Carbon gain and water use in pinyon pine-juniper woodlands of northern New Mexico: field versus phytotron chamber measure- ments. Tree Physiology. 9:59-67. 118 Lauenroth, w.K.; Urban, D.L.; Coffin, D.P; Parton, W.J.; Shugart, H.; Kirchner, IB.; and Smith, T.M. 1993. Modeling vegetation structure — ecosystem process interactions across sites and ecosystems. Ecological Modeling. 67:49-80. Lin, T.C.; Rich, EM.; Heisler, D.A.; and Barnes, F.J. 1992. Influences of canopy geometry on near-ground solar radiation and water balances on pinyon-juniper and ponderosa pine woodlands. American Society for Photogrammetry and Remote Sensing, 1992 Annual Meeting. 285-294. Los Alamos National Laboratory. 1994. Installation Work Plan for Environmental Restoration. LA-UR-93-3987. Nyhan, J.W.; Hacker, L.W.; Calhoun ,T.E.; and Young, D.L. 1978. Los Alamos National Laboratory, Los Alamos, New Mexico, 102. Rich, PM.; Hughes, G.S.; Barnes, FJ. 1993. Using GIS to reconstruct canopy architecture and model ecological processes in pinyon-juniper woodlands. Proceedings of the Thirteenth Annual ESRI User Conference. Vol- ume 2:435-445. Richie, J.T. 1972. A model for predicting evaporation from a row crop with incomplete cover. Water Re- sources Research. 8(5): 1204-1213. Roundy, B.A.; Blackburn, W.H.; and Eckert, R.E. Jr. 1978. Influence of prescribed burning on infiltration and sediment production in pinyon-juniper woodland, Nevada. Journal of Range Management. 31(4) :250- 253. Schlesinger, W.H.; Reynolds, J.F.; Cunningham, G.L.; Huenneke, L.F.; Jarrell, W.M.; Virginia, R.A.; and Whitford, W.G. 1990. Biological feedbacks in global desertification. Science. 247: 1043-1048. Seyfried, M.S.; Wilcox, B.P 1995. Scale and the nature of spatial variability: field examples having implications for hydrologic modeling. Water Resources Research, (in press). Seyfried, M.S. 1991. Infiltration patterns from simulated rainfall on a semiarid rangeland soil. Soil Science So- ciety of America Journal. 55:1726-1734. Walter, H. 1971. Ecology of tropical and subtropical vegetation. Oliver and Boyd, Edinburgh. Whittaker, R. H. 1975. Communities and Ecosystems. Macmillan, New York. Wilcox, B.P 1994. Runoff and erosion in intercanopy zones of pinyon-juniper woodlands. Journal of Range Management. 47:285-295. Wilcox, B.P; Seyfried, M.S.; Cooley, K.R.; and Hanson, C.L. 1991. Runoff characteristics of sagebrush rangelands: modeling implications. Journal of Soil and Water Conservation. 46(2): 153- 158. Wilcox, B.P; Wood, M.K.; Tromble, J.M. 1988. Factors influencing infiltrability of semiarid mountain slopes. Journal of Range Management. 41(3): 197-206. Wilcox, B.P; Wood, M.K. 1989. Factors influencing in- terrill erosion from semiarid slopes in New Mexico. Journal of Range Management. 42(l):66-70. Yeakley, J.A.; Moen, R.A.; Breshears, D.D.; and Nungesser, M.K. 1994. Response of North Ameri- can ecosystem models to multi-annual periodicities in temperature and precipitation. Landscape Ecol- ogy, (in press). 119 Understory Production and Composition in Pinon-juniper Woodlands in New Mexico Rex D. Pieper1 Abstract. — Herbaceous standing crop was determined on three locations within pinon-juniper woodlands on the Fort Stanton Research Station in the Sacramento Mountains near Capitan, NM for a 12 year period. End-of- season standing crop varied from over 1000 kg/ha to a low of less than 200 kg/ha. These differences are comparable to those for other locations in the Southwest. During the period from 1964 to 1975 variations in July-August precipitation explained over 40% of the variation in herbaceous standing crop. Across 25 pinon-juniper stands on Fort Stanton, overstory canopy cover was negatively related to total understory herbage weight. A 2nd polynomial equation explained over 93% of the variation in understory herbage weight attributed to overstory cover while two C3 species Pipto- chaetium fimbriatum and Stipa neomexicana, were positively related to can- opy cover. Two-way cabling resulted in increased herbaceous standing crop of more than 185 kg/ha over a 6-year period after cabling on four loca- tion on the research station. Other studies indicated varying responses to tree reduction. Manipulating overstory cover appears to be one way of al- tering understory production and composition. Pinon-juniper vegetation re- sponds to topographic variation, but the response is quite variable across these southwestern woodlands. INTRODUCTION Pinon-juniper woodlands serve as important habitat for wildlife and domestic livestock in the Southwest. These woodlands occupy over 12.5 million hectares in Arizona and New Mexico alone (Dortignac 1960 and West and others 1975). In some cases these woodlands occur as extensive stands while on other cases, they occur as scattered and disjunct elements. Initially these woodlands were valued mainly as habitat for animals, but re- cently other resource values have been recognized within the pinon-juniper woodland (Bledsoe and Fowler 1992 and Fowler and others 1985). These other products include fence posts, fuelwood, poles, wildings, pihon nuts, and Christmas trees. Yet most of these woodlands continue to be used for livestock grazing and wildlife habitat. Under- story components are especially important for live- stock, but an understanding of their relationship to other environmental variables is necessary for a 1 Professor of Range Science, Department of Animal and Range Sciences, New Mexico State University, Las Cruces, NM. more complete understanding of the ecology of these woodlands. Summaries of information concerning floris- tics of understory vegetation on pinon-juniper woodlands have been presented by Pieper (1977) and West and others (1975). While juniper and pinon-juniper communities in the West are rep- resented by only 9 or 10 tree species (Tausch, this volume, Tueller and Clark 1975), the understory component is much more varied. West and oth- ers (1975) listed over 30 species of shrubs, 8 suc- culent species, 30 grass species, and 14 fob species on 20 locations in the western U.S. In northern portions of the woodlands understory grasses are dominated by C3 cool-season species while C4 warm-season grassers are more impor- tant in southern and southeastern areas (West and others 1975). Pieper (1977) compared un- derstory vegetation for areas in northern regions with that from the southwestern and southeast- ern regions. 120 The objective of this paper is to compare un- derstory vegetation in pinon-juniper woodlands in New Mexico and the Southwest and to discuss fac- tors influencing development of understory vege- tation. Table 2. - Simple linear correlation coefficients between depend- ent variables of total herbage, blue grama, and forb biomass and precipitation for certain periods of the year at Fort Stan- ton, NM (from Pieper and others 1971). Ann. Ppt Dependent variables Oct-Sept J J AS1 JAS JA July Aug Total biomass 0.47* 0.68** 0.68** 0.69** 0.51* 0.65** HERBAGE PRODUCTION Clary (1978) summarized results of understory herbaceous biomass for 6 different studies in dif- ferent locations. These data ranged from less than 50 kg/ha before removal of overstory trees to more than 2200 kg/ha following tree removal. In addition to large spatial variations in herbage biomass, there is considerable variation from year to year. At Fort Stanton in the foothills of the Sacramento Moun- tains in south-central New Mexico, end-of-season biomass varied from less than 200 kg/ha to over 1000 kg/ha at three locations over a 12 year period (Table 1). There were two severe short-term drought periods during this period in 1970-71 and in 1974 (Pieper an Donart 1973, Pieper and others 1992). In general growing conditions were more fa- vorable during the 1960s than during the 1970s and herbage production shows a general decline from the peak in 1967 to 1975 (Table 1). FACTORS INFLUENCING UNDERSTORY BIOMASS Understory biomass is a function of several abi- otic as well as biotic factors. Some of these relation- ships have been studied in detail and others only superficially. Quantitative relationships have been developed for the influence of precipitation and overstory canopy on herbaceous biomass. Table 1. - Herbage standing crop at three locations with pirion- juniper woodlands in the Sacramento Mountains in southcen- tral New Mexico (mean ± standard error). Locations Year 1 2 3 kg/ha 1964 486 ±35 435 ±44 336 ±30 1965 592 ±50 595±47 660 ±44 1966 570 ±65 400±22 630 ±52 1967 725 ±52 899 ±60 1048±47 1968 529 ±33 556 ±57 543 ±27 1969 462 ±37 391 ±44 527±33 1970 543 ±35 667 ±55 585 ±67 1971 335 ±25 250 ±22 375±35 1972 285±35 225±18 330±44 1973 440±40 385 ±39 410 ±50 1974 185±15 235 ±28 300 ±32 1975 330 ±40 385 ±35 425 ±38 Average 457 452 514 Blue grama biomass 0.37 0.51* 0.55* 0.54* 0.44 0.46* Forb biomass 0.36 0.46* 0.24 0.29 0.14 0.33 June, July, August, and September. *P < 0.05 **P < 0.01 Precipitation Several combinations of monthly precipitation were related to total, blue grama, and forb biomass at Fort Stanton in south-central New Mexico using a combination of location and yearly data sets (Pieper an others 1971). Summer precipitation was significantly related to total herbage biomass with 47% of the variation in understory herbaceous biomass determined by differences in July and August precipitation (Table 2). The regression equation was Y = 123.4 + 3.2X where Y = herbage biomass in kg/ha and X = July and August precipitation in milliliters. August precipitation was the best single month predictor of total herbaceous biomass. Correlation coeffi- cients were lower for blue grama and forb biomass. The only significant relationship for forb biomass was with June, July, August, and September pre- cipitation (Table 2). These data were obtained from mature pinon-juniper woodland stands on Fort Stanton and stand density and overstory variation was minimal. Clary (1971) reported that different plant groups responded to precipitation distributions in different manners. Perennial grass biomass was controlled largely by winter-summer precipitation while annual grasses responded to summer pre- cipitation. Shrubs were related more closely to winter precipitation than that received during other periods. Tree Overstory Several studies have evaluated the relationship between overstory cover or tree basal area and un- derstory production or cover (Pieper 1983). These relationships are similar although the functional form of the relationship varies among locations and 121 Table 3. - Equations depicting relationships between herbage production and overstory cover in southwestern pinon juniper Dependent Independent Location Variable Variable(s) Equation Author(s) Northern Herbage Percent Y=597-527(-e*'36x),/2 Jameson 1967 Arizona Biomass Canopy cover Northern (Utah Herbage Basal Area (Xi) Y= 345.4 =129.4X! Clary and Arizona Juniper) Biomass June-Aug. PPT (X3) + 28.3X3 others 1 974 Northern (Alligator Herbage Basal Area (X^ Y=593.3=357.0X1 Clary and Arizona Juniper) Biomass June-Aug. PPT (X2) + 75.4X2 others Great Basin Understory Overstory Y=20.36-0.752X Tausch and cover cover others 1 981 New Mexico Herbage Overstory Y=1 062-1 70X+0.82X2 Pieper 1 990 years. Herbage biomass is reduced by small incre- ments of canopy in dense stands but at lower can- opy cover, the curve is much flatter. Several studies report equations for prediction of understory bio- mass from overstory cover (Table 3) (Clary 1974, Jameson 1967, Pieper 1990, and Tausch and others 1981) while others showed the curves without equations (Arnold and others 1964, Dalen and Snyder 1987, and Short and others 1977). Reduc- tion of canopy cover from 60 to 70% increases her- baceous biomass by small amounts, but reduction from 20 to 10% increases herbage biomass consid- erably. The only linear relationship was that re- ported by Tausch and others (1981) for several areas throughout the Great Basin but the depend- ent variable in their study was herbaceous cover and not biomass. Large variations in understory response to tree reductions is to be expected since the studies represent different areas with different soils, climatic patterns, and time since treatment. Arnold and others (1964) reported that maximum response in understory production occurred 10 to 12 years following treatment. Bledsoe and Fowler (1992) found that herbage response was highest 4 to 7 years following cutting on two sites and sooner on another site in New Mexico. Degree of distur- bance and precipitation patterns following control undoubtedly influence the response of understory species. While the general pattern is for herbaceous biomass to decrease as overstory cover increases, some species actually increase as canopy cover in- creases. Clary and Morrison (1973) reported that cool-season grasses such as Poa fendleriana, Sitanion hystrix, Koeleria cristata, and Agropyron smithii were more abundant under the canopy of large alligator junipers (Juniperus deyrpeana) than in the open spaces in northern New Mexico. In New Mexico, several cool season grasses were positively corre- lated with overstory tree canopy cover (Armentrout and Pieper 1988, Pieper, 1991, and Schott and Pieper 1985). Table 4. - Percent increase in understory herbaceous biomass following pinon-juniper control in the Southwest. Location % Increase Authors Arizona 1227 Clary and Jameson 1981 Arizona 339 Clary 1971 Arizona 105 O'Rourke and Ogden 1969 Nevada 441 Everett and Sharrow 1 985 New Mexico Bledsoe and Fowler 1 992 Taos 316 Chama 1617 Cloudcroft 2907 New Mexico 33 Rippell and Others 1983 The strong relationship between overstory canopy cover and herbaceous biomass suggests that reduction of trees in the pinon-juniper woodlands would increase forage for livestock and possibly game. Consequently, during the 1950s and 1960s, large scale woodland control projects were undertaken on public lands (Aro 1971). Many of these projects were not evalu- ated, but several studies have been conducted to determine the response of herbaceous vegetation to pinon-juniper reductions. The results of these studies have been varied (Table 4) with under- story biomass increasing form less than 35% to nearly 30 times the untreated control. In Arizona O'Rourke and Ogden (1969) reported understory vegetation did not respond to pinon-juniper control on some sites because of high calcium carbonate content of the soil and probable low phosphorus availability. The overall response in herbaceous biomass was higher following con- trol on three locations and nearly double on the four locations studied (Table 4). Table 5 shows that herbaceous biomass increased over 185 kg/ha during a six-year period following two way cabling at Fort Stanton. The increase was greater than 200 kg/ha for most years, but the average was reduced by the lack of response during the second year following the cabling. 122 Table 5. - Herbaceous biomass (kg/ha) on four locations on Fort Stanton following two way cabling for tree control. Year Control Cabled Difference 1975 600 797 +197 1976 642 375 -267 1977 556 859 +303 1978 455 685 +230 1979 512 825 +313 1980 530 717 +187 Topography Considerable research has been conducted on topographic influences on tree species in the pi- non-juniper woodland, but little on understory re- sponses. Lymbery and Pieper (1983) reported that several species were found on most locations sam- pled in south-central New Mexico, and were not restricted to a particular topographic position or aspect. In western New Mexico, cool season grasses tended to be more abundant at higher, mesic habi- tats while warm season grasses were more abun- dant at lower elevations (Hill 1990). In Nevada, Everett and Others (1983) reported that understory species responded differently to the presence of one-seed juniper (Juniperus monosperma) depending on the aspect (north, west, or south). Shrubby components appear more responsive to topographic variables than herbaceous species. Density of cholla cactus was greater on relatively dry south facing slopes compared to north slopes in pifion-juniper woodlands in the foothills of the Sacramento mountains in south-central New Mex- ico. Other shrubs were not greatly influenced by topographic position. Mountain mahogany (Cercocarpus moujntanus) density was greater at higher and intermediate elevations in western New Mexico while grey oak (Quercus grisea) was more abundant at intermediate elevations (2000 to 2200 m). Other species occurred on different as- pects and elevations. LITERATURE CITED Armentrout, Susan M.; Rex D. Pieper. 1988. Plant distri- bution surrounding Rocky Mountain pinyon pine and oneseed juniper in south-central New Mexico. Journal of Range Management. 41:139-143. Arnold, Joseph E; Jameson, Donald A.; Reid, Elbert H. 1964. The pinyon-juniper type of Arizona: Effects of grazing, fire, and tree control. U.S. Department of Agriculture Production Research Report 84. Aro, Richard S. 1971. Elevation on pinyon-juniper con- version to grassland. Journal of Range Management. 24:188-197. Bledsoe, F. Neal; Fowler, John M. 1992. Economic evaluation of the forage-fiber response to pinyon- juniper thinning. New Mexico State University Agri- cultural Experiment Station Bulletin 753. Clary, Warren P 1971. Effects of Utah juniper removal on herbage yields from Springerville soils. Journal of Range Management. 24:373-378. Clary Warren P; Baker, Malchus B. Jr; O'Connell, Paul E; Johnsen, Thomas N.; Campbell, Ralph N. 1974. Ef- fects of pinyon-juniper removal on natural resource products and uses in Arizona. Res. Pap. RM-128. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain forest and Range Experi- ment Station. Clary, Warren P; Morrison, Douglas C. 1973. Large alliga- tor junipers benefit early-spring forage. Journal of Range Management. 26:70-71. Clary, Warren P; Jameson, Donald A. 1981. Herbage production following tree and shrub removal in the pinyon-juniper type of Arizona. Journal of Range Management: 34-109-113. Clary, Warren P 1987. Herbage production and live- stock grazing on pinyon-juniper woodlands. In: Gen. Tech. Rep. INT-215. U.S. Department of Agriculture, forest Service, Intermountain Forest and Range Ex- periment Station. Ogden, UT. Dalen, Raymond S.; Snyder, William R. 1987. Economic and social aspects of pinyon-juniper treatment — then and now. Gen. Tech. Rep. INT-215. Ogden, UT: U.S. Department of Agriculture, Forest Service, Inter- mountain Forest and Range Experiment Station. Dortignac, E. J. 1960. Water yield from pinyon-juniper woodland. In: Water yield in relation to environment in the southwestern United States. Southwestern and Rocky Mountain Division of the American Asso- ciation for the Advancement of Science. Alpine, TX. Everett, Richard L.; Sharrow, Steven H. 1985. Response of grass species to tree harvesting in singleleaf pin- yon-Utah juniper stands. Res. Rap. INT-334. U.S. Department of Agriculture, Forest Service, Inter- mountain Forest and Range Experiment Station. Everett, Richard L.; Sharrow, Steven H.; Meeuwig, Rich- ard O. 1983. Pinyon-juniper woodland understory distribution patterns and species associations. Bulle- tin of the Torrey Botanical club. 110:454-463. Fowler, John M.; Peacock, Bruce E.; Schaber, Michael J. 1985. Pinyon-juniper woodland type in New Mexico: asses or liability. New Mexico State University Agri- cultural Experiment Station Bulletin 718. Hill, Alison. 1990. Ecology and classification of the pin- yon-juniper woodlands in western New Mexico. Ph.D. Dissertation, New Mexico State University, Las Cruces, NM. Jameson, Donald A. 1967. The relationship of tree over- story and herbaceous understory vegetation. Journal of Range Management. 20:247-249. 123 Lymbery, Gordon A.; Pieper, Rex D. 1983. Ecology of pinyon-juniper vegetation in the northern Sacra- mento Mountains, New Mexico State University Ag- ricultural Experiment Station Bulletin 698. O'Rourke, James T.; Ogden, Phil R. 1969. Vegetative re- sponse following pinyon-juniper control in Arizona. Journal of Range Management. 26:302-303. Pieper, Rex D.; Montoya, James R.; Groce, V. Lynn. 1971. Site characteristics on pinyon-juniper and blue grama ranges in south-central New Mexico. New Mexico State University Agricultural Experiment Station Bulletin 573. Pieper, Rex D. 1977. The southwestern pinyon-juniper ecosystem. In: Gen. Tech. Rep. RM-39. U.S. Depart- ment of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. Fort Collins, CO. Pieper, Rex D. 1983. Overstory-understory relationships; pinyon-juniper woodlands. In: Overstory-understory relationships in western forests. Western Regional Publications 1. Colorado Agricultural Experiment Station, Fort Collins, CO. Pieper, Rex D. 1990. Overstory-understory relations in pinyon-juniper woodlands in New Mexico. Journal of Range Management. 43:413-415. Pieper, Rex D.; Donart, Gary B. 1973. Drought effects on blue grama rangelands. New Mexico State University Livestock Feeders Report. Las Cruces, NM. Pieper, Rex D.; Parker, E. E; Donart, Gary B.; Wallace, Joe D.; Wright, Jimmie D. 1992. Cattle and vegetational response to 4-pasture and continuous grazing sys- tems. New Mexico State University Agricultural Ex- periment Station Bulletin 756. Schott, Martin R.; Pieper, Rex D. 1985. Influence of can- opy characteristics of one-seed juniper on understory grasses. Journal of Range Management. 38:328-331. Short, Henry L.; Evans, Wain; Boeker, Erwin L. 1977. The use of natural and modified pinyon-juniper wood- lands by deer and elk. Journal of Wildlife Manage- ment. 41:543-559. Tausch, Robin J.; West, Neil E.; Nabi, A.A. 1981. Tree age and dominance patterns in Great Basin pinyon- juniper woodlands. Journal of Range Management. 34:259-264. Tueller, Paul T; Clark, James. 1975. Autecology of pinyon- juniper species of the Great Basin. In: The pinyon- juniper ecosystem: A symposium. Utah State Uni- versity, Logan, UT. West, Neil E.; Rea, Kenneth L.; Tausch, Robin J. 1975. Ba- sic synecological relationships in juniper-pinyon woodlands. In: The pinyon-juniper ecosystem: A symposium. Utah State University, Logan, UT. 124 A Checklist for Ecosystem Management in Southwestern Pinon-Juniper Earl F. Aldon1, Reggie Fletcher2, and Doug Shaw2 Abstract. — It has been estimated that 3.5 million acres of pinon-juniper in the Southwestern Region are in unsatisfactory watershed condition, the current Pihon-Juniper Initiative in the Region places priority on restoration of these acres in conjunction with the Forest Service's move toward ecosys- tem management. Using a checklist format, this paper lists important eco- logical aspects useful in developing and implementing desired future conditions designed to move degraded areas toward a more sustainable condition. While the needs of each site are different, utilization of this checklist will ensure projects implemented have a broad focus based on scientific information. References are provided for additional in-depth re- view for many items while others relay on the author's experience. The checklist can be utilized as an aid for implementation of the IRM process. INTRODUCTION The current emphasis on ecosystem man- agement has resulted in much talk and writing on the philosophy and principles of ecosystem management from both the physical/biological and human dimension. A team of scientists, chartered by the Regional Forester, recently pub- lished "An Ecological Basis for Ecosystem Man- agement" in which six guiding principles are stated. The premise for sustaining ecosystems and protecting biodiversity now and into the fu- ture is to manage ecosystems such that structure, composition, and function of all elements includ- ing their frequency, distribution, and natural ex- tinction, are conserved. (Kaufmann et al. 1994) Another Regional team has published a bro- chure in which eight human dimension princi- ples and strategies are discussed (R-3 Human Dimension Study Group). The philosophies and principles outlined in both these reports provide an important basis for policy decisions. Few rules or side boards are available to guide specific efforts in ecosystem management at particular forest locations for the Pihon- Juniper (P-J) type. An ecosystem principles filter 1 Retired, Rocky Mountain Forest and Range Experiment Station, Albuquerque, NM. USDA Forest Service, Southwestern Region, Albuquerque, NM. 125 is described in the Kaufmann report. Economic and social needs can be tested against a filter of physical and biological principles. Also, eco- nomic and ecological needs can be tested against a filter of human dimension principles. The con- cept of using niters is good and clear, but how do you do it? This checklist is an attempt to provide some in- gredients for the filter and move toward ecosystem thinking based on principles before embarking on a specific project in Southwestern Pihon-Juniper ecosystems. The current Southwestern Regional initiative for the P-J type is focused on restoration ecology and requires guidelines to insure man- agement strategies are based on biological, physical and social science. By following this checklist in planning one should be fairly certain major omis- sions will be avoided. The checklist is based on re- search information and where possible referenced. It is not meant to be complete and final, but an evolving checjdist that will require periodic updat- ing as more and better information becomes avail- able. It assumes all applicable laws (Endangered Species Act, National Forest Management Act, Fed- eral Land Policy Management Act, National Envi- ronmental Policy Act, etc.) will be followed. An ecosystem needs assessment proposed by Kaufmann, et al. is a useful way to organize this checklist. The assessment consists of: (1) Defining the analysis area. (2) Finding and describing refer- ence conditions. (3) Describing and understanding existing conditions. (4) Applying a course filter analysis. (5) Applying a fine filter analysis if neces- sary. (6) Describing ecosystem needs and capabili- ties. CHECKLIST Defining the Analysis Area • Begin by looking over a broad scale (for example a province or section on the ecological scale) to see how a particular proposed project area fits in a larger eco- logical prospective. • Human dimension and physi- cal/biological considerations extend be- yond jurisdictional boundaries. Effects of actions at a particular site can be far reaching and cumulative. • Identify the stakeholders at this level of scale? What are their desires for the fu- ture? • What is the potential biological diversity? • What are the economic factors that need to be considered at the larger scale? • Focus in on the particular project area where restoration ecology appears needed (watershed, range allotment, etc.). • Determine the priority of the site specific proposal in the context of the broader scale assessment and the predicted sus- tainability deficiencies surfaced by other disciplines. • Is priority given to the 3.5 million acres of National Forests in the Southwest in un- satisfactory soil and watershed condition? (Spann 1993) (Shaw 1993) • Is priority given to areas of social or cul- tural need? • Is priority given to threatened, endan- gered or sensitive species habitat needs? • Is priority given to heritage resource sites? • What are the key indicators of system health or sustainability at the broader scale? REFERENCE CONDITIONS • Are there undisturbed examples of the same or similar ecosystems available for direct evaluation of natural ecosystem structure, composition, and function? • Has literature been searched for informa- tion on historic and/or reference condi- tions? • Has the nature, frequency, intensity, and scale of disturbance been identified and considered at the local and landscape scale? • Has an attempt been made to identify the natural biological diversity of reference conditions and of the project area? • Have changes in climate, soils, hydrology, and human use patterns been considered in deciding the appropriate understand- ing and use of reference conditions? COURSE AND FINE FILTER ANALYSIS • Will planned actions over time result in an array of vegetative structure and com- position aggregates similar to that under which the project area developed? • Will planned actions effect critical habitat features for T.E.&S. species? EXISTING CONDITIONS • Have state water quality reports been consulted to determine water quality standards and water quality status rela- tive to meeting beneficial uses? • Have stakeholders been consulted about past, current and future use of the area? • Have field investigations been conducted to assess canopy structure and composi- tion changes that have occurred over time to understand factors of change (drought, 126 fire, ecosystem dynamics)? (Betancourt et al. 1993) • Have you referred to historical records for the area - aerial photos, range allotment plans, old cultural treatments (chaining, pushing, farming, etc.) to understand causative factors of change. • Have ecosystem dynamics such as fire and drought been considered in under- standing changes? (Cook et al. 1991) • Have you decided what P-J habitat types are represented in the project area? An uneven aged mature pinon stand requires much different management than a juni- per savannah. (USDA - Forest Service 1987) (Dick-Peddie 1993) • Has Terrestrial Ecosystem Survey data been used, if available to further define habitat? (USDA - Forest Service variable dates) • Have collaborative partnerships been sought among stakeholders? (Giusti 1993) (USDA - Forest Service 1993) (USDA - For- est Service Southwestern Region 1993) • Are public involvement recommendations for desired future condition biologically feasible? (Garcia 1993) • Do proposed project areas include sacred places for American Indians or other cul- tures? Are archeological sites noted? (Koyiyumptewa 1993, Cartledge et al. 1993, Miller etal. 1993) • What is the status of key indicators of sys- tem health or sustainability at the broader scale? ECOSYSTEM NEEDS AND CAPABILITIES • Will plans protect or improve soil quality? • Will state water quality standards be met? Are Best Management Practices pre- scribed? • Will plans result in a visually desirable mosaic on the landscape? • Will riparian areas be protected or im- proved? • Will heritage resources be protected? • Will threatened, endangered, and sensi- tive species habitats be protected? • Will actions improve or enhance the pub- lics understanding or appreciation of the P-J ecosystem? • Will human life style needs be protected? • Can existing human uses be sustained? • Is sufficient flexibility in human use pro- vided to weather drought years. • Are there potential commercial areas for pihon nut production? (Cunningham et al. 1993) • Can areas be managed for nut production within the project area? (Norwick et al. 1993, Cunningham et al. 1993) • Can Christmas trees be harvested from the area? (Barger and Ffolliott, 1972) • Are there areas needing revegetation with grasses? • Are there remnant native herbaceous or shrubby seed sources that can, when managed, expand native species? (Scholl et al. 1986, Johnson, T. N. Jr. 1987) • Are grass species present in mosaic pat- terns rather than continuous one-species stands? Have lA acre openings been pro- vided to insure a positive blue gramma response? • Where herbaceous species are growing directly under tree canopies (little-seed or pinon ricegrass) have % acre clumps of trees been retained to prevent desicca- tion? • Are treatment sites needed as a corridor or barrier to movement of animal species? • Are climatic cycles considered in schedul- ing actions? (Betancourt 1993) • Have cryptogram cover been considered in management plans? (Ladyman et al. 1993) • Are insects and diseases in the stand in- ventoried and included in management plans? Are mistletoe infected trees scheduled for removal? (Rogers 1993) (Shaw, C. G. et al. 1994) (Gottfried, G. J. et al. 1994) 127 Are there areas of bitterbrush and cliffrose present that could be enhanced through management for winter range for wild- life? (Suminski 1993) Is dead and down material present and could it be enhanced by management for nutrient enhancement and erosion pro- tection? (Ernest et al. 1993) Do fuelwood harvest plans consider making small openings, utilizing smaller size trees, lopping and scattering pifion limbs to prevent slash buildup? Can harvest plans be outlined in steps to prevent desiccation of understory vegeta- tion? Will slash placement provide erosion pro- tection and increase organic soil content? Do pinon seedlings need protected with slash to enhance survival? Is protection from grazing for at least two grazing seasons possible? If fire is considered for management has nutrient depletion as a result of fire been considered? (Perry 1993) If fire is an appropriate tool to move us toward a desired condition, is herbaceous recovery sufficient to control erosion and survive the fire? Can fire prescriptions be followed while meeting other resource objectives? (Wright 1982) Are follow-up maintenance needs and funding planned? Are your monitoring goals purposeful and retrievable from a corporate data base? (Brady unpublished) Are provisions made to monitor, store, and retrieve understory vegetation changes? Are plans made to monitor key indicators of system health or sustainability at the broader scales? Considering everything contemplated are we sure we will do no harm? LITERATURE CITED Many papers referred to in this checklist ap- pear in either of the following 2 references. Rather than repeat these citations the paper will be cited as a I or II to refer to its location. I. Aldon, Earl E; Shaw, Douglas W, technical co- ordinators. 1993. Managing Pinyon-Juniper Ecosystems for Sustainability and Social Needs; proceedings of the symposium 1993 April 26-30; Santa Fe, New Mexico. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agricul- ture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 169 p. II. Everett, Richard L., compiler. Proceedings — pinon-juniper conference; 1986 January 13-16; Reno, NV. General Technical Report INT-215. Ogden, UT: U.S. Department of Agriculture, For- est Service, Intermountain Research Station; 1987. 581 p. 1. Brady, Ward, John Cook, Earl F. Aldon, unpublished. COVER - A decision support system for ecological monitoring. Arizona State University Dept. of Ag. - Business, Tempe, AZ. 2. Perry, Hazel. 1993. Soil nutrient research on the He- ber Ranger District Apache-Sitgreaves National For- est, p 149-152. IN:I 3. Shaw, D.W. 1993. Pinyon-Juniper initiative in the Southwestern Region p 12-13. IN: I. 4. Scholl, David G. and Earl F. Aldon. 1986. Grass es- tablishment on uranium exploration sites in New Mexico, pp. 95-98. In: Proceedings of the National meeting of the American Society for Surface Mining and Reclamation (Jackson, MS., March 17-20), 1986. p. 219. Harper, Jarvis and Bill Ploss eds. American Soci- ety for Surface Mining and Reclamation, 21 Grand- view Dr., Princeton, WV 24740. 5. Wright, Henry A., and Arthur W. Bailey. 1982. Fire ecology. Ch. 9, p. 195-208. 6. Spann, Charles L. 1993. Procedural guidelines for de- veloping soil and water conservation practices in Pinyon-Juniper ecosystems, p. 159-161. IN. I. 7. Suminski, Rita R. 1993. Management implications for mule deer winter range in Northern Pinyon-Juniper. p. 133-139. IN:I. 8. Rogers, Terrence J. 1993. Insect and disease associ- ates of the Pinyon-Juniper woodlands, p. 124-125. IN:I. 128 9. Ernest, K.A., Earl E Aldon, and E. Muldavin. 1993. Woody debris in undisturbed Pinyon-Juniper wood- lands of New Mexico, p. 117-123. IN: I. 10. Ladyman, J.A., E. Muldavin, and Reggie Fletcher. 1993. Pattern and relationships of terrestrial crypto- gram cover in two Pinyon-Juniper communities in New Mexico, p. 97-104. IN: I. 11. Betancourt, Julio L. et al. 1993. Influence of history and climate on New Mexico Pinyon-Juniper wood- lands p. 42-73. IN: I. 12. Dick-Peddie, William A. 1993. New Mexico vegeta- tion: past, present and future. University of New Mexico Press. Albuquerque, NM p. 244. 13. USDA Forest Service - Southwestern Region. Various dates. Terrestrial Ecosystem Report (TES) Albuquer- que, NM. 14. USDA Forest Service - Southwestern Region. 1987. Forest and woodland habitat types of Arizona and New Mexico. 3 vol. Albuquerque, NM. 15. USDA Forest Service - Southwestern Region. 1993. Integrated Resource Management. The road to eco- system management. 4th ed. Albuquerque, NM. p 28. 16. USDA Forest Service. 1993. The power of collabora- tive planning. Report of the National Workshop. Washington, D.C. p. 12. 17. Giusti, Gregory A. 1993. Model for the conservation of biological diversity through bio-regional planning. Proceedings Society of American Foresters annual meeting. Indianapolis, Indiana. 1993. Society of American Foresters, Bethesda, MD 20814-2198. 18. Cook, John W; Brady, Ward W; Aldon, Earl F. 1991. Handbook for converting Parker loop frequency data to basal area. Gen. Tech. Rep. RM-212. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Sta- tion. 22 p. 19. Shaw, C. G., F. G. Hawksworth, D. Bennett, G. Sanchez-Martinez, and B. M. Tkacz. 1994. Diseases and insects of pine and there implications for sustain- ability in forests of the Southwestern United States and Northern Mexico, p. 36-50. In: Manzanilla, Hugo; Shaw, Douglas; Aguirre-Bravo, Celedonio; Iglesias Gutierrez, Leonel; Hamre, R. H., tech. coords. 1993. Making Sustainability Operational: Fourth Mexico/U.S. Symposium; 1993 April 19-23; Santa Fe, NM. Gen. Tech. Rep. RM-240. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 232 P- 20. Gottfried, Gerald J, and Peter F. Ffolliott. 1994. Silvi- cultural prescriptions for sustained productivity of the Southwestern Pinyon-Juniper and encinal woodlands, p. 185-192. In: Manzanilla, Hugo; Shaw, Douglas; Aguirre-Bravo, Celedonio; Iglesias Gutier- rez, Leonel; Hamre, R. H., tech. coords. 1993. Mak- ing Sustainability Operational: Fourth Mexico/U.S. Symposium; 1993 April 19-23; Santa Fe, NM. Gen. Tech. Rep. RM-240. Fort Collins, CO: U.S. Depart- ment of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 232 p. 21. Garcia, Maria Teresa. 1993. Traditional use of pinyon- juniper woodland resources, p. 79-81. IN: I. 22. Koyiyumptewa, Bruce K. 1993. Spiritual values of the pinyon-juniper woodland: A Hopi's perspective, p. 19-20. IN: I. 23. Cartledge, T. R., and Judith G. Propper. 1993. Pin- yon-juniper ecosystems through time: Information and insights from the past. p. 63-71. IN: I 24. Miller, Ronald K. and Steven K. Albert. 1993. Zuni cultural relationships to pinyon-juniper woodlands, p. 74-78. IN: I. 25. Cunningham, Gary, Jim Fisher, and John Mexal. 1993. Establishing research, management, and harvest ar- eas for pinyon nut production, p. 85-88. IN: I. 26. Norwick, Jim, Dennis Garcia, and Bill Torgensen. 1993. Commercial leases and permits for pinyon nut harvesting, p. 24-25. IN: I. 27. Barger, Roland L, and Peter F. Ffoliott. 1972. Physical characteristics and utilization of major woodland tree species in Arizona. USDA Forest Service, Research Paper RM-83, 80 p. Rocky Mountain Forest and Range Experiment Station. Fort Collins, CO. 28. Johnson, Thomas N. Jr. 1987. Seeding pinyon- juniper sites in the Southwest. IN: II. 29. Kauffmann, M.R., R.TGraham, D.A.Boyce Jr., W.H. Moir, L.Perry, R.T.Reynolds, R.L.Bassett, PMehlhop, C.B.Edminster, W.M.Block and PS.Corn. 1994. An Ecological Basis For Ecosystem Management. U.S. DA. Forest Service, General Technical Report RM-246. Rocky Mountain Forest and Range Experi- ment Station. Fort Collins, Colorado 80526. 30. R-3 and Rocky Mountain Forest and Range Experi- ment Station Human Dimension Study Group. 1994. The Human Dimension in Sustainable Ecosystem Management. USDA Forest Service. Albuquerque, NM 87102. 129 Woodland Inventory Procedures and Analyses Conducted for Management Planning Purposes on Indian Lands John Waconda1 INTRODUCTION The Bureau of Indian Affairs (B.I.A), Albuquer- que Area Office (A.A.O.), Branch of Forestry main- tains the responsibility of managing approximately 1.2 million acres of tribally owned woodland forest resources on twenty-four separate Indian reserva- tions in Southern Colorado and New Mexico. Initiatives undertaken by Indian tribes and the B.I.A. in the 1980's spawned an increased aware- ness of Indian woodlands management culminat- ing in a woodlands study report produced in 1988, and later led to special woodlands management funding beginning in 1989. Fortunately, Congres- sional woodland management funding has contin- ued to be appropriated on a year-to-year basis since then. This special funding has been utilized to fund various types of woodland management projects nationwide including inventory data collection and inventory analyses. The 1988 B.I.A. woodlands study addressed several deficiencies associated with tribal wood- lands management, two specifically related to woodland resource inventories. The first was that specific volume information was lacking for all woodland types and ownerships; and secondly, substantial woodland acreages failed to be in- cluded in current forest management plans, and that harvesting and land management activities were generally done without silvicultural direction or environmental clearance (Bureau of Indian Affairs 1988). Reasons for these deficiencies within the B.I.A. comparable to other land management agencies may be attributed to the lack of funding specifically earmarked for woodlands manage- ment due to past emphasis placed on commercial timber management. Traditionally the woodland resource was perceived of as having little if any commercial value and accordingly received 1 Woodland Management Forester, Bureau of Indian Affairs, Albu- querque Area Office. minimal attention. Nevertheless, over the years public attitudes have changed and have resulted in transforming land management policies and practices; whereby the woodland ecosystem has received more attention ranging from preserva- tion to increased development initiatives. Regardless of management direction, sound in- ventory data is essential for determining forested areas, evaluating land productivity and forest health, estimating forest volumes, and to measure forest trends toward meeting desired management objectives. Inventory data collection and associated inventory analyses procedures are just two steps within the overall B.I.A. Forest Management Inven- tory and Management Planning initiative. Other steps include remote sensing, mapping, environ- mental analyses, and developing forest manage- ment plans. Presently, the B.I.A - A.A.O. has completed inventory data collection procedures on thirteen reservations, and completed the associated in- ventory analyses for three of those reservations (fig. 1). Planned for execution within the next three years are an additional nine more reserva- tion woodland inventories. To accomplish this task the A.A.O. has employed two full-time in- ventory foresters located at various field loca- tions to conduct the inventory work on a pro- jected schedule. The purpose of this paper is to describe stan- dard A.A.O. planning, implementation, and data analysis procedures used in previous woodland inventory projects. These procedures were de- veloped to meet the needs of the B.I.A. as well as individual tribes, and may be unique depending on particular circumstances. Also addressed are several unusual situations encountered on these projects and the related project design modifica- tions needed to deal with these special circum- stances. 130 TRIBES WITH INVENTORY DATA COLLECTED 1 . Acoma Pueblo 2. Isleta Pueblo 3. Jemez Pueblo 4. Laguna Pueblo 5. Mescalero Apache 6. Nambe Pueblo 7. Ramah Navajo 8. San lldefonso Pueblo 9. Santa Clara Pueblo 10. Southern Ute 1 1 . Taos Pueblo 12. Ute Mountian Ute 13. Zuni Pueblo TRIBES WITH COMPLETED INVENTORY ANALYSES 4. Laguna Pueblo 6. Nambe Pueblo 10. Southern Ute Figure 1. — Bureau of Indian Appairs, Albuquerque Area Office woodland inventory locations. INVENTORY PLANNING AND DESIGN To facilitate program definition and manage- ment priorities, the B.I. A. employs a national cate- gorization system that prioritizes reservations and/or Indian properties based on current forest land data. Indian reservations/properties are clas- sified into a five category system based on forest resource acreages and regional resource value standards (Table 1). This system is then used as a basis to determine the inventory design and mini- mum accuracy standards for a particular project. Although, due to the extreme forest resource vari- ability among Indian nations within the B.I.A. jurisdiction, regional B.I.A. Area Directors may amend inventory standards to accommodate unique local situations. Presently, the A.A.O. has implemented and/or planned inventory projects on tribal lands classified as Category 1, 2, 3, and 4. Planning Planning cannot be over emphasized in the forest inventory process. Quite often, planning does not receive the consideration and time that it deserves; instead, efforts are sometimes focused on collecting field data and analytical phases of the project. When this happens, inventories can be inefficient and, in some cases may not serve man- agement objectives. In the B.I.A. we are fortunate to have a more decentralized forest management system relative to the more familiar forest man- agement agencies. Usually the B.I.A. foresters that plan an inventory are the same people who will implement it, and then analyze and use its data. Consequently, the B.I.A. Forest manager often assumes the role of the planner, data collector, and decision maker for the complete inventory and management planning process. 131 Table 1. -Bureau of Indian Affairs Reservation/Indian Property Forest Resource Categorization System. Category Forest Resource Description Inventory Standards Major Forested Reservation - Comprised of more than 100,000 acres of commercial timberland in trust, or more than 1 .0 MMBM harvest of timber product annually. Based on a 10 yr. reoccurring permanent plot inventory subject to an accuracy standard of 5% error at 1 standard deviation on both basal area and the primary unit of volume measurement used in the calculation of allowable cut. Minor Forested Reservation - Comprised of less than 10,000 acres of commercial timberland in trust, and less than 1 .0 MMBM harvest of timber products annually, or whose forest resource is determined by the Area Office to be of significant commercial timber value. Significant Woodland Reservation - Comprised of an identifiable forest area of any size which is lacking a timberland component, and whose forest resource is determined by the Area Office to be of significant commercial woodland value. Minimally Forested Reservation - Comprised of an identifiable forest area of any size determined by the Area Office to be of minor commercial value at this time. Reservation or Indian property with forest land that the Bureau is charged with some degree of legal respon- sibility, but the land is not in trust status. Based on a 10 yr. reoccurring temporary plot inventory supported by permanent growth plots, as deemed necessary, and subject to an accuracy standard of 1 0% error at 1 standard deviation for basal area and/or the primary unit of volume measurement used in the calculation of the allow- able cut. Based on a 10 yr. reoccurring stand reconnais- sance inventory, or as may be more restrictively defined by the Area Office. An inventory may be required more frequently due to the occurrence of environmental and/or developmental impacts. Based on a 10 yr. reoccurring stand reconnais- sance inventory, or as may be required more frequently do to the occurrence of environmental and/or developmental impacts. Based on Area Office determination. A distinctive process to B.I.A. forest inventory planning is the inclusion of tribal participation. Tribes are encouraged to not only participate in the inventory planning initiative but, also in the entire forest management planning process. This is im- portant because often times tribes may desire to make changes that will affect their commercial woodlands land base such as: establishing condi- tional or total woodland reserve areas, protect religious and cultural significant sites, construct housing developments, develop recreation areas, and maintain roadless areas, etc.. Cases of active tribal participation include working with tribes that are interested in develop- ing or expanding woodland enterprises on their reservations. Inventory information for these tribes may be specialized to include information about the resources utilized in these enterprises. Many tribes may restrict harvests of live green trees and only allow tribal members to harvest dead and down material for fuelwood. Therefore, may re- quest that the inventory include information on the availability and volume of dead sound fuel- wood material. These are just a few examples of active tribal participation incorporated into the B.I.A. woodland inventory planning process. The most important step in planning a new inventory, or in remeasuring an existing system, is to identify its objectives. The planner must first identify what forest resource information will be needed to effectively manage the woodland resource over a planned management period. In almost all instances, cost is of primary importance in planning B.I.A. woodland inventories. The objective is to obtain reliable estimates to a predetermined precision standard for the lowest possible expenditure. Once the needs and objectives for woodland management have been identified, the inventory then becomes a sampling problem. Design Standard woodland inventory project design planning includes three main steps: 1) obtaining necessary remote sensing imagery and related information; 2) preparing specialized forest maps or map overlays; and 3) selecting a sampling method and design that addresses data collection requirements. On large timbered reservations, Continuous Forest Inventory (CFI) systems are used to provide the information needed to prepare and modify forest management plans. However, on reserva- tions with lesser timber resources, inventories using temporary installed plots can be used and growth information is obtained from increment 132 cores. Ordinarily, CFI systems are not used on these reservations due to the high cost of perma- nent plot installation, maintenance, and remeas- urement in comparison with overall timber values and timber sale income. Planning for a one-time inventory using temporarily installed plots, is essentially the same as any inventory for manage- ment planning except a commitment does not have to be made for plot maintenance and remeasure- ment, this greatly reduces overall costs. Within the A.A.O., woodland inventories have consisted of both CFI and temporary plot systems. As such, the designs have generally remained the same for either system, based on a systematic sampling method using fixed area plots established by developing a systematic grid. Grid size is usu- ally based on specific forest acreages and the strata that will be inventoried. Commonly the standard grid system is developed using the Universal Transverse Mercator System (UTM) overlaid on the desired sampling strata, utilizing the most current USGS 1:24,000 scale, 7.5 minute quads as base maps. Stratification is usually done using cover type maps delineated on the most recent aerial photog- raphy. However, B.I.A. woodland projects often lack this information due to the absence of previ- ous forest management planning on woodland habitat. In most cases, accurate, current, and thor- ough reservation-wide woodland resource typing is limited and therefore, area stratification is often refined while the field inventory is implemented. In cases where reservations may be lacking woodland resource typing data, woodland inven- tories are based on stratification done using exist- ing forest resource data in the Geographic Infor- mation System (GIS). For example, commercial woodland stands may be determined using crown closure percentages or density classifications, a slope limitation factor, and its accessibility to an existing road system. This information may be available in the GIS, and not on aerial photos or cover type maps. Essentially, the standard within the A.A.O. is to implement low cost inventory planning practices often using innovative design procedures while ensuring a quality product. PLOT LAYOUT AND DATA COLLECTION The current woodland inventory plot design used in A.A.O. projects is a single-unit, fixed area, circular plot (fig. 2). The plot contains two subdivi- sions superimposed within the major plot to collect additional selected data. Based on the typical woodland forest conditions a 1/20 acre major plot size has been determined to be the most effective. On this major plot, all timber species 5" DBH and greater and woodland species 3" DRC and greater are measured as well as mortality trees. Minor plot 1 is composed of the northeast quadrant of the major plot (l/80th of an acre). In addition to the measurement procedures con- Plot Center Minor Plot 1 NE Quadrant (1/80 acre) Minor Plot 2 1/100 acre (11.8' Radius) Major Plot 1/20 acre (26.3' Radius) Figure 2.— Woodland inventory plot design. 133 ducted on the major plot, trees within this minor plot are bored to determine radial growth and age, with the exception of the junipers. Minor plot 2 is circular in shape, 1/100 acre in size. This minor plot is superimposed over the major plot center. This is a regeneration plot in which timber tree species <5" DBH and woodland species <3" DRC are tallied. For our inventories, establishing plot locations in the field is basically the same whether a tempo- rary inventory and permanent inventory system is used, with only minor modifications between the two. In permanent inventories, a location header form is completed by the installation crew. This information will be used later to locate the plot when remeasurement is initiated. A procedure recently employed on permanent inventories in the Albuquerque Area is the use of a Global Posi- tioning System (GPS) to document permanent plot locations. This information is then incorporated into the existing GIS data base for later referencing. The inventory data collection system used by the A.A.O. is a modified version of the U.S. Forest Service, Region 3, Stage II Stand Exam. Modifica- tions to this basic system were made to accurately measure woodland species and to obtain informa- tion as identified in the planning process. After the permanent plot center is established and the perimeter of the plots identified on the ground, the marking of the tally trees begins. Each live woodland species tree larger than 3" DRC on the plot is identified and tagged with an aluminum tag numbered sequentially in a clockwise direction from the north. The tag is secured to the base of the tree below stump height and facing the plot center. This point is used because accurate identification of the trees that may be harvested between measure- ments is important. An aluminum nail is driven into each measured tree J/4 inch below the point of DRC measurement. Since the nail marks the point for all future diameter measurements, it is impor- tant that it be positioned accurately. Measuring woodland tree species is not as clearcut as measuring timber species. Junipers, pinons, and oaks in mis region often grow with multi-stems and are extremely variable in form. Often special attention is necessary when determining accurate diameter measurement points due to butt swelling and stem forking. When stems are clumped together and appear to be from the same origin with a unified crown, they are treated as a single trees. Accordingly, an equivalent diameter at root collar (EDRC) must be calculated. The EDRC is equal to the square root of the sum of all qualifying squared stem diameters. The apparent differences in form, growth, regen- eration, causes of mortality, management objectives, and forest products between timber species and woodland species undoubtedly result in differences in inventory procedures. The following is a brief discussion related to some of the major factors that should be considered when conducting woodlands inventories; growth and tree form influence diameter measurements, deterrnining what a stem is as op- posed to a branch is subjective. Realistic guidelines must be developed to insure consistency in stem measurements. Form damage such as sweeps, crooks, severe lean, and forks are not recorded, due to the lack of effect on major woodland forest products. Although, tree heights must take into account com- mon sweeps and horizontal growth. Measuring accurate woodland growth is a diffi- cult process, and may not be possible if relying on tree bore-backs exclusively. On A.A.O. woodland inventory projects, woodland growth is based on increment cores obtained solely from pirion trees for two reasons. First, junipers are free growers whose rings do not correspond to annual growth, especially in drought situations. Secondly, junipers are very difficult and almost impossible to bore because of its wood structure. In any case obtaining and reading increment cores from pinons are diffi- cult and much more time consuming. Consequently, current growth estimates are based on pihon growth rates that are applied to all woodland tree species. A more accurate estimate of forest growth can only be assessed after the next inventory measurement is conducted which will then appraise actual growth of all species. Another important aspect in A.A.O. woodland inventories is the collection of mortality tree infor- mation, basically for fuelwood management pur- poses. On most Southwestern reservations wood is i still the preferred cooking and heating source. i Accordingly, fuelwood harvesting is an important forest activity for many tribal members, and on some reservations fuelwood enterprises have been developed creating a demand and income for woodcutters. Tribes have thus begun to recognize the importance of this resource and in response have initiated fuelwood harvesting policies and L regulations. Some traditional tribal policies have included: restricting harvests to only tribal mem- jj bers, establishing limits on the amount of fuelwood that may be cut per household, and prohibiting the ,v cutting of live green trees thus allowing only har- vests of dead and down fuelwood. Clearly with fuelwood collection being such an important aspect of tribal lifestyle, and with the Bureau being responsible for managing tribal 134 woodlands under a sustained yield principle, fuelwood inventory data is extremely important and essential. Mortality assessment includes determining if the tree has died in the last five years. Trees dead within the last five years usually still retain a ma- jority of their branches and bark, and sometimes foliage. In addition, inventory crews informally determine in the field if a dead sound tree meets local fuelwood specifications related to its size and soundness. If the tree is suitable for fuelwood, diameter and length measurements are recorded to compute volumes, as well as percentage of bark remaining intact. This will give forest managers an estimate of current dead fuelwood available, while the next measurement will yield and accurate estimate of annual accumulation and projected use of dead fuelwood. Woodland insect and disease problems are unique and often times are difficult to identify much less discern without training and experience. Mistletoes, Arceuthobium divarication on pihon and Phoradendron juniperinun on juniper are common, and in the case of the pihon variety, it is oftentimes difficult to detect. Interestingly, the mistletoe sometimes seems to transcend into a dormant stage or inactive period where the fruiting bodies disap- pear. In the initial stages of conducting woodland in- ventories, A.A.O. foresters spent considerable time developing and refining standardized procedures. Since men an area-wide inventory handbook has been developed that outlines standard procedures. This has proven to be invaluable especially when training inventory personnel. Although standard procedures have been developed for B.I.A.- A.A.O. woodland inventories, it is safe to assume that modifications and fme-tuning may be necessary when initiating other projects. lands and incorporating tribal participation in the planning process. Assuring sustained yields of multi-use forest resources is also an important overriding management objective in the B.I.A.. Therefore, major emphasis is devoted to adminis- tering sound resource management planning. An essential component of the planning proc- ess is a resource management inventory, whereby a determination within a reasonable degree of accu- racy may be made of volumes, growth, condition, health, and forest trends. Inventory data will be analyzed, discussed, and presented in terms that can be incorporated into forest management plans. Inventory analyses are crucial to the manage- ment planning process. The analysis report should describe the sampling method and procedures, describe current forest conditions, summarize stocking and growth information, include compari- sons to previous inventories, analyze forest trends, and determine allowable harvesting levels. Depending on forest resource acreages and values, the complexity of the inventory analyses differ. Reservations with vast forest resources will undoubtedly require an inventory analysis with a higher degree of intricacy. Another important factor is that tribal participation is essential during this step. Input from the tribe is required when final land classification procedures require desig- nation of management units, and identifying cul- turally significant areas warranting special atten- tion. The goal is to have an inventory analysis that is useful for management planning purposes. Whether a particular tribes' desire is to preserve or develop its woodland resource, a thorough and complete inventory should provide the informa- tion to assist the tribe in making an educated deci- sion. CONCLUSION Indian tribes in the Southwest have become in- creasingly interested in woodland resource man- agement. Appropriately, the B.I.A. is responsible for implementing prudent forest management prac- tices in accordance with state-of-the-art standards, while recognizing the unique character of Indian LITERATURE CITED Bureau of Indian Affairs. 1988. Native American Wood- land Resources: A National Overview, Assessing the Resource Potential and Management Needs. USDI, Bureau of Indian Affairs, Branch of Forest Resources Planning. 135 Watershed Restoration Through Integrated Resource Management on Public and Private Rangelands Sid Goodloe1 Abstract. — Until recently much of the rangeland in the western United States was in a serious downward trend. Water quality and quantity were declining as the result of the continuous livestock grazing practices em- ployed at the turn of the century followed by 80 years of fire suppression. Thirty-five years of integrated/holistic resource management at the Carrizo Valley Ranch site has reversed this trend. In addition to restoration of rangeland productivity, the riparian area on the ranch has been restored, wildlife populations enhanced and perennial streamflow restored. The practical experience gained at the ranch should be useful to private land- owners, public land managers, and water quality agencies throughout the brittle ecosystems of the southwestern United States. Some of the tech- niques used at Carrizo Valley Ranch are being demonstrated on an adja- cent watershed in the Smokey Bear Ranger District of the Lincoln National Forest. BACKGROUND The shortgrass rangelands found in the west- ern United States are generally harsh ecosystems. Careful management of these areas is essential if they are to maintain sustained production or re- cover from past land management mistakes (Stoddart et al. 1975). Many watersheds in the west contribute massive loads of sediment washed from the land surface or scoured from eroding gullies and streambanks to the streams and rivers which drain them. The New Mexico Environment De- partment reports that 95% of the state's surface water is impacted by nonpoint source pollution (NMED 1990) and that turbidity is one of the major causes of use impairment in these waters (NMED 1988). Reports by early surveyors, naturalists and trappers detail the abundance of grass and clear, clean water found on these same watersheds (Leopold 1933/1991), a sharp contrast to the condi- tions seen today. Many factors have contributed to the drastic changes that can be seen in the rangeland watersheds of the western United States, but most range management professionals agree that the 1 0wner/Operator Carrizo Valley Ranch, Box 598, Capitan, NM 88316 heavy stocking rates and the continuous grazing practiced at the end of the 1800's followed by increasingly efficient fire suppression are the leading causes of these changes. H. L. Bently and E. O. Wooten, early agricultural agents in Texas and New Mexico, described the situation: "In a short time every acre of grass was stocked beyond its fullest capacity . . . The grasses were entirely consumed, the very roots were trampled into the dust and destroyed" (Bently 1898). "The stockman could not protect the range from himself, because any improvement of his range was only an inducement for someone else to bring stock in upon it; so he put the extra stock on himself" (Wooten 1908). As a result, native grasses were replaced by sagebrush, mesquite, juniper, and other invading brush species that were less suited for holding soil in place (Chaney et al. 1990) and which were more efficient at water extraction (Stoddart et al. 1975). Topsoil, which requires thousands of years to develop in harsh ecosystems washed away; gullies formed from unchecked, concentrated runoff; streambanks eroded and downcut; water tables lowered; and perennial streams became intermittent or dry (Chaney et al. 1990, Platts 1990). 136 The ability of the land to recover from these ef- fects has been greatly reduced because the entire ecosystem had been so radically altered. The harshness of the environment contributes to the difficulty in reestablishing the climax or the highest ecological condition of the range. As a result, sim- ple manipulation of a single range management factor, such as reducing livestock numbers, is not sufficient to result in significant environmental improvement (DeBano and Schmidt 1989). These systems will take many years to recover by them- selves. Direct actions aimed at total watershed rehabilitation and applied in a holistic and inte- grated system are necessary to ensure the restora- tion of western watersheds and associated natural resources of water, timber, grass, wildlife and fish- eries (Platts 1990). This type of integrated or holis- tic resource management has been successfully demonstrated on the Carrizo Valley Ranch. INTEGRATED RESOURCE MANAGEMENT ON PRIVATE LANDS There are many definitions of Inte- grated/Holistic Resource Management (IRM), but I like to define it as the integration of all compo- nents, economic, human and environmental, into a synergistic, comprehensive plan that allows man- agement for long-term sustainability rather than short-term production. This type of management is essential for protecting valuable natural re- sources found in our western watersheds and is also an essential management tool for protecting the entire planet. Considering the unlimited sup- ply of examples of bad natural resource manage- ment in every state of the US and in every country in the world, it is clear that we are now charged with the responsibility of not only managing the resources under our jurisdiction in an integrated manner, but we must also inform politicians and populations everywhere that we are no longer in the pioneering/unplanned development mode. We have reached the point that resource interrelation- ships must be recognized and development planned accordingly. Pressing needs of growing populations must be met, but not at the expense of the ecosystem's sustainability. INITIAL ACTIONS My ranch is located in the South Central Moun- tains of New Mexico at about 7000 feet elevation. Average annual precipitation is about 46 cm (18 inches), one half of which falls as snow. The soils range from gravelly hillsides to clay and clay loam bottoms. Watercourses on the ranch were actively eroded and brush infestation flourishing when I purchased the property. My most demanding problem was the homogeneous vegetative compo- sition and low herbage production. The major grass found was an almost pure, tightly packed turf of bluegramma that grew very little because of its sod-bound condition. A major portion of the ranch had scattered to thick stands of pihon-juniper that were even-aged populations. Areas between the trees as well as directly under the canopy were bare and subject to erosion. I began to study the origination of this eroded, brush-infested condition. I realized that year-long grazing and brush infestation were severely limit- ing herbage production. My initial strategies were 1) to divide the ranch into summer and winter pastures so I could at least reserve some winter grazing and 2) to begin a systematic brush control program. Although these changes were beneficial, it was not until I spent time in Rhodesia (Zimbabwe) in 1964 that I experienced first hand and began to understand the principles of Short Duration Grazing in action and the dynamics of an open savannah ecosystem. I recorded my findings in a paper published in the November 1969 issue of the Journal of Range Management, returned to my ranch, and, after some very low budget fencing, put these principles into practice. ROTATIONAL GRAZING SYSTEM I divided large paddocks into much smaller ones using posts cut on the ranch to support a three-wire suspension fence. Paddock division was planned according to topography, existing fences and available water — not in the wagon wheel or grazing cell pattern often advocated. Once the rotation had become established, the cattle practi- cally moved themselves — anticipating paddock changes. I found that graze and rest periods could be adjusted to fit the current precipitation and season of use. I also found that as the vegetative growth rate increases, so should the frequency of rotation and that rotation during the dormant season was not necessary. My initial goal now became "to produce the maximum pounds of mar- ketable beef per hectare while improving range condition." This naive, but commendable, goal was economically impractical in a period of low beef prices, so I needed to find other profitable uses of available resources. 137 ADDITIONAL INCOME SOURCE Fee hunting of deer and turkey became a sig- nificant income producer immediately after I built a cabin to facilitate game harvest. As a result, im- proved wildlife habitat and overall aesthetic qual- ity became my secondary goal. RETURN TO CLIMAX CONDITION AS THE PRIMARY GOAL The pieces of the puzzle then began to fall into place. I realized that if fish and beaver appeared on the 600-year-old Indian petroglyphs on my place, there certainly must have been running streams where I now found only arroyos with steep banks and dry rocky bottoms. I researched 100-year-old surveyor's notes that described the terrain as an open savannah rather than an almost solid canopy of invading brush species. I realized that the invading brush, made possible by year- round grazing and 80 years of total fire suppres- sion, was not only removing most of the moisture from the soil, but was also shutting down herbage growth, thereby causing sheet and gully erosion. I recognized that although I had previously dis- counted a return to climax or near-climax condi- tion, I might be able to make economic sense out of that approach if it became my primary goal. I visualized the open savannah as it was over 100 years ago, with mixed conifers on the north slopes and the highly productive riparian areas that made up the mosaic of the Carrizo Valley. BRUSH MANAGEMENT AND WATERSHED STABILIZATION I then began to implement a cautious return to climax in a manner that was economically justifi- able in my situation. Mechanical removal of invad- ing pihon-juniper in an area that requires 10 to 15 hectares per animal unit could not be justified because costs were higher than land values. How- ever, some mechanical brush control in the better soil types was required, as was erosion control (i.e., reseeding, pushing invading brush into active gullies and building water retention dams). It was necessary to finance this using other available resources. Selective thinning of young invaders, followed by prescribed burning and reseeding with native grass species became the major thrust of the plan to return to a climax ecosystem. The by-products — fence posts, fuel wood, vigas, trees for landscaping, and Christmas trees — financed the plan. Another beneficial by-product was the increase in mule deer population, not only because of habitat improve- ment, but because ponderosa pine vigas must be cut and peeled during the winter months. This provided an adequate supply of green browse (tree tops) throughout the winter, resulting in a signifi- cant (30 to 50%) increase in the fawn crop. The open savannah created contains 500 to 800 year old juniper trees, scattered ponderosa pines, and is carpeted with a mix of warm and cool season grasses and forbs. I have found that because deer and turkey evolved under this type of ecosystem, they seem to prefer it to the contiguous, brush- infested public land. This is what I call an "eco- recreation benefit/' These factors sharply increased income from hunting and paid for more of the necessary mechanical rehabilitation work. THE ROLE OF FIRE The long sought after open savannah is now well established in the Carrizo Valley, but it must be maintained with periodic fire, as it was in the climax. Tree ring research in New Mexico indicates that most forest areas burned, on the average, at 7 to 10 year intervals (Stoddart et al. 1975) before fire suppression began. The key to the successful maintenance burn is the fuel load (as well as the climatic conditions, of course). There must be enough herbaceous material to carry a fire which is hot enough to kill brush but cool enough not to damage the beneficial species. The damaging fires in Yellowstone a few years ago demonstrated that the no-burn policy, which originated in the ecol- ogically different European forests, was an incor- rect choice for western watersheds. Now after many years of fire suppression, similar fuel loading is evident throughout the Western United States, which makes the initial prescribed burn risky, to say the least. LIVESTOCK SUITED TO THEIR ENVIRONMENT The pivotal economic component of my opera- tion is the production of weaner calves, both for breeding and beef. Low-input, sustained produc- tion is my goal and is achieved by using an animal that is fine tuned to the environment and that produces a desirable, marketable product. The hostile factors in our environment are snow, cold, 138 wind and dry weather. A cow that can produce under these conditions must be, first of all, fertile in that environment. She should be black so that wind and snow will not cause or aggravate pink eye and cancer eye. Black, of course, absorbs as much sparse winter sunlight as is possible and black udders do not blister in spring snow storms. The animal that fulfills all these requirements is a composite breed that I have developed through 20 years of selective breeding called the Alpine Black — three-quarters Angus and one-quarter beef- type Brown Swiss. Just as the Zebu composites fit the Gulf Coast and southern deserts, the Alpine Black fits the western mountains of Northern America. TANGIBLE BENEFITS The road back to climax has revealed many changes in 30 years. Water sources that were dry now have permanent running water and lush riparian areas. Grass production has increased dramatically and provided more carrying capacity. Alpine black cattle are in sync with their environ- ment and their habitat has improved as well. Recreation potential is greatly enhanced due to a more pleasing aesthetic atmosphere and larger wildlife populations. APPLICABILITY OF CASE STUDY RESULTS TO WESTERN WATERSHEDS The pifion-juniper (PJ) complex comprises more than 63 million acres of the rangeland in the southwest. This ecotype is a critical component of the arid region. Pifion and juniper generally form the intermediary boundary between the flatter grassland type climax community found on the lower slopes and the conifer forest climax commu- nity of the mountain tops. Considerable debate regarding the density of the PJ canopy in climax conditions has hindered some watershed restora- tion efforts. Most range conservationists agree, however, that much of the PJ found on the lower slopes has escaped it's original range and modified some of the original savannah type ecosystem to a more woodland type. Originally the PJ occupied a discrete ecotone in many watersheds, but lack of fire and overuse by livestock have left these once stable areas in poor condition. Many, however, have a high potential for range improvement and revegetation. In areas where the PJ complex is in especially poor condition, range improvement can substantially reduce the erosion and sedimentation originating from these degraded areas (Stoddart et al. 1975). Some of the most informed members of the environmental community support restoration of western watersheds, but question the removal of pinon and juniper vegetation from those areas where the species are in the climax community. As opposed to brush removal and range reseeding on areas historically known or reasoned to be grass- land, brush removal on certain areas can have the potential to increase sedimentation and erosion rather than decrease it. Information gained from the Carrizo Valley Ranch can be useful to managers needing to determine if brush management efforts can be reasonably and safely completed and a sustainable system established. Riparian areas and the water they surround are of especial consequence in arid ecosystems. These areas constitute only about 2% of the total western acreage, yet they are among the most productive and valuable lands. DeBano and Schmidt (1989) have described the relationship of upland water- shed condition to riparian condition and found, not surprisingly, a direct correlation between de- graded upland watershed condition and degraded riparian area condition. They concluded that adequate treatment of all critical areas in the upper watershed is necessary to provide a stable and sustainable riparian area and is critical when at- tempting any riparian restoration project. On Carrizo Valley Ranch, we completed most of the upper watershed work (stabilizing gullies, remov- ing invading brush, and revegetating bare ground) before being able to maintain a stable riparian area. Chaney and his coworkers (1990) and Platts (1990) found that maintenance of riparian areas, once restored, requires a different grazing strategy than upland sites. Although I have done some riparian corridor fencing which works to protect the ripar- ian area from livestock access, I have demonstrated that as long as the principle of limited and man- aged access is applied, fencing is not always a required component. The key to the effective riparian protection demonstrated at Carrizo Valley Ranch is protection during the growing season if possible and rapid rotation when not. ECOSYSTEM MANAGEMENT ON PUBLIC LANDS - CARRIZO DEMONSTRATION AREA The watershed above the Carrizo Valley Ranch is part of the Smokey Bear Ranger District of the Lincoln National Forest. In 1989 the USFS began a watershed restoration and demonstration project 139 on 55,000 acres of National Forest and private land as part of its ecological approach to multiple-use management of the PJ ecosystem. The project area contains large expanses of continuous canopy pinon-juniper that prior to the introduction of livestock in the 1800's and subsequent fire sup- pression supported a wide variety of native grass plants. As the range degraded, trees out-competed grass for available moisture and soon much of the productive soil beneath these dense woodland stands eroded away, leaving behind an extensive and active gully system that continues to transport silt-laden water into streams and rivers (Edwards, 1990). With the urging of private land owners, who for years had to contend with the deposition of millions of tons of sediment that originated on National Forest land and who had demonstrated that complete watershed rehabilitation was not possible on their private landholdings, a unique, cooperative, watershed-based project was begun. The project focuses on soil stabilization practices, vegetation management, water resource develop- ment, vehicular travel management and sound grazing management practices. The project's goals include control of soil erosion, stabilization of steep gully slopes, restoration of permanent riparian vegetation and the rehabilitation of native rangelands to support a sustainable mix of native grass and woody plants. As the result of treatments begun in 1989, cool season native species of grass and forbs long absent from the National Forest have returned; in several drainages, springs have begun to flow again; and a wide variety of upland and riparian wildlife spe- cies have returned to the area to make use of the increased edge areas, water supplies and addi- tional food sources. On private lands adjacent to the forest, benefits have also been reported. In one area, 4,800 cubic yards of sediment from gully and sheet erosion originating on National Forest land were cleaned out of a pond. The following spring, after implementation of watershed restoration treatments on the forest, a spring which had not run for 50 years began to flow and continued to flow throughout the summer, filling the pond with clear water. The pond has now been stocked with trout and catfish. SUMMARY Integrated resource management is the profes- sional vernacular describing what managers do who are in tune with efficient, sustained use of the resources that are their responsibility. If the use of one resource affects the health or production of another adversely, then the whole is diminished and economic and environmental costs are guaran- teed to surface somewhere, sometime. Common sense and vision provide the foundation for bringing all parts of the whole together into a comprehensive management plan. Interestingly enough, as are many things in life, it is elusive because it is so simple. And yet, if we intend long term survival, we must implement this approach in every phase of Natural Resource Management. As watershed restoration and rehabilitation work continues, it is important to understand that there will never be sufficient government resources to treat every problem in every area. Thus, success lies in demonstrating techniques such as those developed on Carrizo Valley Ranch which provide internal and self-sustaining motivation for adop- tion on both private and public lands. REFERENCES Bently, H. L. 1898. "Grasses and Forage Plants of Central Texas" Bulletin No. 10. USDA Special Agent in Charge of Grass Experiments, Abilene TX. Chaney, E., Elmore, W, Platts, W. S. 1990. Livestock Graz- ing on Western Riparian Area, US Environmental Pro- tection Agency, Denver, CO. DeBano, L. D., Schmidt, L. J. (1989). Improving Southwest- ern Riparian Areas Through Watershed Management USDA Forest Service General Technical Report RM- 182, Rocky Mountain Forest and Range Experiment Station, Ft. Collins, CO. Edwards, R. 1991. Carrizo Demonstration Area, USFS. Lincoln National Forest, Smokey Bear Ranger Dis- trict. USFS Goodloe, S. 1990. "Twenty Years of Integrated/Holistic Resource Management" from Integrated Resource Management Symposium, Morelia, Mexico, March 27, 1990. Leopold, A. 1993/1991. "The Virgin Southwest" re- printed in The River of the Mother of God and Other Es- says by Aldo Leopold. Flander, S., Baird, J., eds., Uni- versity of Wisconsin Press 1991. New Mexico Environment Department 1990. Biennial Water Quality Report. NMED, Santa Fe, NM. Platts, W. S. (1990). Managing Fisheries and Wildlife on Rangelands Grazed by Livestock. Nevada Department of Wildlife, December 1990. Stoddart, L. A., Smith, A. D., and Box, T. W. 1975. Range Management, Third edition. McGraw Hill Book Com- pany, St. Louis, MO. Wooten, E. O. 1908. "The Range Problem in New Mex- ico" in Bulletin #10, Agri. Expt. Station, New Mexico College of Ag. and Mech. Arts. 140 Can't We All Just Get Along" Jon S. Bumstead1 During the past several years I've worked for the Forest Service in many different jobs: as a For- est Planner, University Liaison, Regional Social Sci- ence Coordinator, and as a University Liaison for Land Management Planning. Currently I am working on the Eastside Ecosystem Management Project as the co-lead of the social science staff on the Science Integration Team. This team is assess- ing the ecological health of the Interior and Upper Columbia River Basin. Lately, there is some debate among science team members questioning "whom we can do sci- ence with," and who can be legitimately involved in this learning/scientific process. I won't go into that discussion now, but it's something to consider while reading this article. In the midst of these Forest Service experiences, and during the time of "New Perspectives," I be- came a realtor and spent many of my off hours moonlighting for a successful real estate company. While working in Flagstaff, I went to night school and obtained a Master's Degree in Applied Sociol- ogy. I think each of these experiences has given me some unique insights that I would like to share. Like most insights, this one begins with a story. One evening about three years ago, in a class on the development and structure of sociological the- ory, I stumbled across a sociological proposition by Dr. Max Weber, one of the founding fathers of social theory. His proposition was, "The more a political authority loses prestige in the external system, the less able it is to remain legitimate." (Zeitlin, 1990). I looked at this statement (which is about equivalent, in sociological terms, to a mathemati- cian saying 2 + 2 = 4), and asked myself if the For- est Service was approaching the point where we would be organized out of existence. Later, in a New Perspectives session in Denver, I was exposed to the work of Dr. Julia Wondolleck, from the Uni- versity of Michigan. In her book, "Public Lands Conflict and Resolution, 1988," she explained how our management paradigm and training experi- ences were focused on a set of values to which the organization steadfastly clings, while American social values have shifted to a less utilitarian view. 1 Forestry Sciences Laboratory, Flagstaff, AZ. Is she right? If so, our organization has been by- passed by a change in basic values. Have we shifted our views enough in the past years to re- gain our legitimacy? A class I took in the development of social theory through qualitative analysis again fo- cused my attention on our operational culture. In this class I learned to observe and record qualitative data. While collecting data, I ob- served that the organization was putting out a lot of messages about "customer service." At the same time, in my personal interactions with a variety of people at all organizational levels, I kept encountering actions that I felt were more self-serving than customer oriented. I kept ask- ing myself why this was happening, despite our good intentions. Finally, as I worked throughout the region in my University Liaison position, I hit on what I be- lieve is the core element that keeps us all from achieving some really worthwhile work It is the failure to communicate effectively both internally and externally. This is probably no surprise but it really is a central, recurring problem. Poor com- munication continually leads to breakdowns in achievement of reasonable solutions to continuing problems and opportunities. I recognize that we all are aware of how difficult it is to communicate effectively. What I have to offer may be new to some of you, and may even turn some of you off. However, if you reflect on my opening questions regarding who it is appropriate to, "do science with," I'll ask you now considering applying some science taken from salespeople. What I have to of- fer today is a selling process that I learned about while I was moonlighting as a realtor. I believe there is much in this process from which we can benefit. The essence of the process is honesty, care- ful listening, and the ability to articulate that you understand how another person feels. Before I describe this process, I would like show you some results from a social survey project I helped construct with the Apache-Sitgreaves Na- tional Forest and the Sociology Department at Northern Arizona University. 141 The management team on the Apache/ Sitgreaves National Forest recognized the importance of understanding people's FEELINGS as a crucial first step in resolving management issues. They asked myself and members of the sociology staff at Northern Arizona University (NAU) to work with them in the construction and analysis of a survey to determine how employees and Forest users felt about the Apache Sitgreaves National Forest timber management program. In my opinion, a lot of Forest Service officials fail to recognize the importance of addressing feelings. We tend to just want the "facts." While the gathering of factual data is an integral part of forest management, the importance of addressing people's feelings and perceptions cannot be ignored (Figure 1). As we worked with the forest staff, we found that some of us were, unintentionally, trying to construct questions that explained the Forest Serv- ice perspective of each question. Dr. Fernandez, a sociology professor from NAU, finally got it through to us that we should seek other people's opinions and perspectives; not explain our own views. Thus, the final questions in the survey were very direct, brief, and nearly free of Forest Service interpretations. I want to show you some of the re- sults, in case you think we've moved past the time of internal and external conflict (Figures 2, 3, 4, 5). We could look at the internal split of opinion and hope it was a sign of organizational balance and health. I want you all to understand, however, that while I was a University Liaison I had people from a variety of ranger districts, national forests, and from within the regional office talk to me in emotionally ridden terms about the "gridlock" and inability of interdisciplinary teams to work to- gether effectively. Further, one of the questions we are currently asking on the Columbia River Basin assessment, asks people whether they trust our agencies' motives and ability to implement ecosys- tem management. With about 50% of the survey responses in, we find that only 30% of the Na- tional sample has moderate or a great deal of trust in the Forest Service. Western Washington and Oregon are at 31%. People living in the Columbia River Basin, currently, are responding at a rate of 33% with moderate or a great deal of trust. An- other sample is of people who have been directly involved and are on our mailing list (over 2000), they have 41% who trust in our abilities to imple- Figure 1. — Dealing it... rather than dealing WITH it. ment ecosystem management but only 29% who have a moderate or great deal of trust in our mo- tives. A previous range land reform survey con- ducted by Dr. Brent Steel from Washington State University found similar responses in a National survey... 32%. If you bat 300 in baseball you are doing pretty good. Does a 30% degree of trust meet our expectations for ourselves? I believe that a first step for each of us is to im- prove our communication skills before we will start to see an upward movement in public trust. With the above demonstrations that a problem does exist, I ask you to consider if there is anything we can learn from the following sales process I learned while moonlighting. The selling process I was taught to use as a real- tor is found in the book "The Best Seller" by Ron Willingham (1984, Prentice Hall). My broker, who had thousands of books in his home, said, "Jon, you only need to read one book on selling and here it is." Willingham's book is rooted in the need to be completely honest, always. Never, never do any- thing that moves you off that center. A second foundation is to find out what the customer needs and wants. Your beliefs and values do not enter the picture until you truly understand the desires and perspective of the customer. 142 100%-/ 90% 80% 70% 60% 50% 40% 30% 20%. 10%- 4 0%M N«137 43% 40% 9% mm 4% 4% SA SO No Opinion Figure 2. — Forest Service employees only: Statement — The Forest Service is doing a good job of managing ecosys- tems. No Opinion Figure 3. — Forest Service employees only: Statement — When timber harvesting is in conflict with recreation activities, priority should be given to recreation activities. While I was forest planner on the Santa Fe National Forest, we frequently planned a conflict resolution strategy prior to appeals negotiations. We would meet before negotiations to determine "our" position and bottom line. In other words, we walked into conflict resolution focused on where we would draw the line in the sand. How open do you think we were to any new ideas? Of course, the appellants were doing the same thing. That was our basic approach to conflict resolution a few years ago: Determine Position — Go Negotiate. In contrast, the sales approach I was taught had these steps: A. Approach I. Interview D. Demonstrate I. Validate* N. Negotiate C. Close 0 Conservationists 23 Forest Business Employees N=113,x2=76, Pr = .0001 5% No Opinion Figure 4. — Conservationists and Forest business employee comparisons: Statement — Recreation activities on the for- ests contribute more to economic stability than do timber harvest activities. 0 Conservationists Forest Business Employees N=lll,x2=78, Pr = .0001 SA SD No Opinion Figure 5. — Conservationists and Forest business employee comparisons: Statement — When timber harvest is in con- flict with recreation activities, priority should be given to recreation activities. The "I" in validate is saying that, "I will take the time to understand and validate your con- cern from your perspective." I still try to apply in my daily work activities. In Willingham's process, negotiation is the next to the last step. In the appeals process we often opened with negotiation discussions. There was hardly a breath of "Hi, how are you" prior to start- ing negotiation. In Willingham's process, a lot of foundation work precedes the negotiation step. During the Approach and Interview steps we were taught to "tune yourself out and tune your customer in." The most important job of successful sales people, according to Willingham, is to listen, listen, listen! The process has a variety of application tech- niques for each phase in the sales process. For in- stance the acronym "F.O.R.M." was used in the Approach phase. The letters stand for Family, Oc- cupation, Recreation, and Message. Most people 143 like to talk about one of these things, and, if you lis- ten carefully, you'll pick up messages from them about what they value most. This process empha- sizes the importance of getting to know your cli- ents on a personal basis. I think this relates closely to what the early district rangers excelled in... "Spittin' and Whittling." Maybe we should think more about going back to the old method. It pays! The positive side of taking the time to get to know each other on a personal basis is extremely impor- tant. In the book, Willingham talks about a sales per- son so committed to the Approach that he would decline trying to potential customers if he was un- able to build a good rapport at the outset. He sim- ply felt people would never buy from him, if he was unable to build an immediate relationship that "felt" right. During the transition from Approach to the Interview phase we were expected to gain a clear understanding of customers needs, desires, and lifestyle priorities. The ability to tune your own values and preferences out is extremely difficult. Try it! Only your customer's perspective is impor- tant during these first phases.. I'll skip to the Validation process next, as I feel this is the area where we could all make the biggest strides in learning to communicate effectively. Willingham insists that it is not enough to just re- peat back what you've heard the customer say. Many of our communication classes suggest we paraphrase the words we have heard from some- one else to demonstrate that we have indeed "heard" them. While that is better than just inter- rupting and stating your opinion, it falls short from Willingham's perspective. He recommends that you personally step in their shoes and validate why their concerns are valid, from their perspec- tive. You can't do this unless you have a clear un- derstanding of their views and can articulate to them that you understand their perspective, not yours\ You should relate their concern to a similar experience you have had with the technique of "Feel, Felt, Found." It goes like this: "I see how you feel about A, B, C. I felt that way too about D, E, F (your similar experience); Here's what I found about D, E, F. By the way, my broker had a Doctorate degree in psychology, and had been a professor at the University of Oklahoma for several years. His brother worked for him as a sales representative and held a Master's degree in sociology. The brother, Dave, used this sales process to perfection. As I watched him work, you could see that he was always totally engaged with his customers. If you watched him for 20-30 minutes he might be in a speaking mode for only five of those minutes. The rest of the time he listened to his customers. He had a knack for keeping dialogue open and flowing. About every week you could count on Dave closing another contract. Some of you may have concerns about this process being used to manipulate people. I know I did. I said to my broker, "Jeez, you guys are dan- gerous!" The broker believed that you couldn't use this process to trick people. He felt that if you tried any of these techniques with anything but complete sincerity and honesty, your customers would read that insincerity and be gone in a min- ute. Further, he stated that by law we were re- quired to reveal everything we knew about a home. Keeping your opinions and feelings out of discussions in the beginning doesn't mean you give up stating your opinion during the negotia- tion and final stages of the sale. Your willingness to learn about and understand your customer's per- spective goes a long way towards building trust. Further, it will make the customer more receptive to your views as you enter the negotiation and closing phases of this process. As he went on, my broker explained how important it is to all of us to know that we have been heard and understood. As I have reflected on this process recently, and done some objective/scientific observations around our work areas, I find that we all (me too!) are so eager to tell people what we think and feel, that we often don't listen carefully enough to gain an under- standing of the other person's perspective. Back to the process, once you've moved through these first four phases of the sales process, it is time to enter into negotiations and, finally, come to a sales agreement. In three years of real es- tate sales I successfully utilized this process only twice (I made other sales, but none felt as good as these two). I am too eager to insert my own opin- ion and talk about myself, and I'm not alone in this weakness. While I was a University Liaison traveling throughout the region, I observed fre- quent failures to achieve complete communication. This was occurring at all levels within the organi- zation. Specifically, the lack of the ability to listen to the point of true understanding was preventing us from doing, or becoming, all that we could and should be. So what can we do? First, we need to bolster our education and training in the human dimen- sion fields, especially vital communication skills. In the long-term, we should continually work at staying involved with citizen groups all throughout the year, not just for the length of each individual 144 project. I know the Prescott National Forest has been actively working on this for several years. Short term recommendations include allowing ample time in our planning processes for meaningful relationship building both within and without the organization. You can't expect to throw a bunch of people in a room and have them automatically function like a team. It takes time! Finally, I think each of us needs to identify and work on our own communication shortcomings. Personally, I noticed that "yeah, but" was maybe the single greatest stopper of effective communica- tion in our organization. It's everywhere! Listen, you'll see. I used to be one of the greatest users of "yeah but." To the best of my knowledge, I have stopped myself from using this phrase. I may still interrupt, but I won't do it with "yeah but." REFERENCES The Best Seller, The New Psychology of Selling and Per- suading People; Willingham Ron; Prentice Hall, 1984. Public Lands Conflict and Resolution, Managing Na- tional Forest Disputes; Wondolleck Julia M.; Plenium Press; 1988- Ideology and the Development of Sociological Theory; Zeitlin Irving M.; Prentice Hall; 1990. WE'RE ALL IN THIS TOGETHER, SO WHAT COMMUNICATION IMPROVEMENTS CAN YOU MAKE? 145 Responding to Tribal Voices in Managing Woodland Resources Ronald K. Miller1 Abstract. — Native American peoples and tribes have utilized and depended on woodland resources for a wide variety of practical and ceremonial uses for centuries. Often, deeply rooted traditional value systems are associated with uses or methods of harvest of the resource. The Bureau of Indian Af- fairs, Division of Forestry, has responded to tribal concerns about woodland ecosystems and is currently seeking to manage these forests in a sustain- able, multi-disciplinary manner guided by tribal culture and tradition. Spe- cific woodland areas are currently being managed for the production of fuelwood, posts and poles, pinon nuts, furniture wood, and even for tradi- tional wood and herbs used in cremation rites. Nonconsumptive uses in- clude woodland areas set aside for religious activities such as ceremonial dances or the planting of prayer sticks, or areas set aside for wildlife habitat or scenic beauty. "Every spring our spiritual elders bless the everlasting prosperity of the pine nuts. The annual Walker River Paiute Pine Nut Festival honors the traditional fall pine nut gathering. Pinon pitch is used to seal our baskets and start camvfires. Pinon wood warms our household during the cold winter months. Pinon pine is essential to the lifestyle and cultural preservation of the Washoe, Northern Paiute, and Shoshone tribes. " — Ben Rupert, Washoe/Northern Paiute "I think, due to past pinon-juniper management practices (chaining projects), some woodlands are depleted of plants, animals, and birds so vital to Hopi spiritual ceremonies. " — Bruce Koyiyumptewa, Hopi "Clearly, the pinon-juniper areas represent a major potential resource for the Pueblos. With unemployment rates exceeding 50% on some Pueblos, the Bureau needs to focus on programs which have the greatest economic potential for the Pueblo people. " — Alvino Lucero, Isleta "The management of our woodlands provide benefits directly and indirectly to tribal members. Directly, by heating our homes, cooking our food, ana by providing income through the sale of firewood. Indirectly, through improved wildlife habitat, healthier woodlands through the control of disease and insects, and improved grazing areas. — Lyman Clayton, San Carlos Apache 1 Woodland Forester, Bureau of Indian Affairs, Phoenix Area Office. 146 INTRODUCTION PAST WOODLAND MANAGEMENT Native American peoples and tribes have util- ized and depended on woodland resources for a wide variety of practical and ceremonial uses for centuries. More importantly, perhaps, most South- western tribes have lived in the pifion-juniper eco- system for centuries. Given that long-term association, the connections between tribes and the resource are often very strong, with deeply rooted traditional value systems associated with uses or methods of harvest of the resource. A natural re- source professional working with a tribe must un- derstand the cultural connections to the resource just as he or she must understand the silvics and ecology of the woodland species being managed. One must also understand that tribal cultures differ and what is acceptable to one tribe is not necessar- ily acceptable to another tribe. The quotes prefacing this paper are from mem- bers of four different Southwestern tribes. They touch on the importance of the pinon-juniper eco- system to tribal people whether that value relates to cultural, spiritual, economic, aesthetic, medici- nal, or subsistence issues. Most likely, the relation- ship with the resource is a combination of all of the above. Often the resource and the culture are so in- tertwined that the various components cannot be separated. Indeed, the Native American world view tends to look at things holistically, rather than as individual components. Tribal relationships with pinon-juniper re- sources have existed for centuries and current lit- erature and ethnobotanical studies expound on these age-old ties (Ackerly 1991, Koyiyumptewa 1993, Miller and Albert 1993, Watahomigie et al. 1982). The Hualapai have even published a beauti- ful, full-color, 20-page booklet in their own lan- guage that discusses their relationship with pinon (Watahomigie et al. 1983). The booklet is entitled Ko, the Hualapai name for pinon. The importance of pifion-juniper woodlands is magnified by their sheer abundance in the South- west. Arizona's woodland area of 14.3 million acres, for instance, is more than 2-1/2 times that of timber- lands (Conner et al. 1990), while Nevada's wood- land (9 million acres) represents almost 12 times that of Nevada's timberlands (Born et al. 1992). Here in Arizona, Indian tribes own 35% of the woodland, making tribes the largest woodland ownership class in the state. Current estimates of Indian-owned woodland in Arizona range from 5.0 million acres (Conner et al. 1990) to 5.6 million acres (Bureau of Indian Affairs 1988). Although the resource is so large, the potential so great, and the cultural ties so deep, integrated management and full consideration of pinon- juniper woodlands is still in its infancy. In the past, woodlands were either ignored, or worse, a war was waged against them in the form of chaining, cabling, pushing, burning, tree crush- ing, grubbing, chopping, and herbiciding. Any- thing, to get rid of those short, scrubby trees. In Arizona alone, nearly one million acres of pinon- juniper woodland were cleared between 1950 and 1959 (Arnold et al. 1964). Another 300,000 acres were cleared between 1960 and 1985 just on Forest Service lands in Arizona and New Mexico (Dalen and Snyder 1987). The Bureau of Land Manage- ment (BLM) was busy as well, chaining half a mil- lion acres of pinon-juniper in the West between 1960 and 1972, 61,000 acres of which were in Ari- zona (Aro 1975). As stated by former Forest Service Southwest Regional Forester, William Hurst (1977), "Perhaps no vegetative type (P-J) has given man so much and been so harshly treated and neglected." The Bureau of Indian Affairs (BIA) was not im- mune to the management strategies of the day. In some places the techniques were different but the intent was the same. On the Hualapai Indian Res- ervation in northwest Arizona, nearly 22,000 acres of pinon-juniper were burned between 1953 and 1963 for conversion purposes (Despain 1987). Meanwhile, on the Fort Apache Reservation, tribal members were kept busy with hand axes clearing 95,000 acres of pifion-juniper by 1958 (Arnold et al. 1964). Chaining also occurred on many southwestern Indian reservations. This is especially ironic since pifion-juniper woodlands contain such an abun- dance of archaeological sites, and chaining proved so destructive to these valuable, nonrenewable re- sources (DeBloois et al. 1975). Irreplaceable tribal history and culture were destroyed in the name of range improvement. McNichols (1987) traced the evolution of P-J management on the Hualapai Res- ervation from no management, to an emphasis on removal or eradication, and finally to a strategy of multiple use and sustained yield. It took an active intervention by tribes themselves, however, to gen- erate the support necessary to establish a national woodlands program. 2The reason given for all this activity was, of course, "range im- provement". However, many of the clearing projects, along with the rea- sons given for them, are seriously questioned by such notable authorities as Dr. Ronald Lanner and Dr. Elbert Little among others (Lanner 1981, 1993; Little 1991, 1993). 147 BACKGROUND On April 15, 1987, Alvino Lucero, Vice President of the Isleta Pueblo Tribal Council, testified before the Bureau of Indian Affairs Agriculture - Range Programs Committee. He presented a strong case detailing the importance of the pinon-juniper woodland resource to the Pueblo people. Primarily focusing on the economic potential of the resource, he stated: "Most of our forests in terms of acreage and wood volumes consist of pinon-juniper trees as opposed to commercial species such as ponderosa pine. For example, the four Pueblos of Acoma, Isleta, Jemez and Zia nave an annual allowable cut of ponderosa pine of about 2 million board feet. This represents combined potential revenues of $20,000 per year. In contrast, these same Pueblos could harvest approximately 1,560 cords of pinon- juniper per year. Based on $100 per cord, this would bring in gross revenues of $156,000 each year. Clearly, the Pinon- juniper areas represent a major potential resource for the Pueblos. With unemployment rates exceeding 50% on some Pueblos, the Bureau needs to focus on programs which have the greatest economic potential for the Pueblo people." The following month, the Southern Pueblos Governors Council wrote a letter to the Assistant Secretary - Indian Affairs on behalf of the nine Southern Pueblos in New Mexico. The Assistant Secretary was asked to expand the BIA's focus and work to include management of woodland re- sources. Within three months, on August 10, 1987, a di- rective was sent from the Assistant Secretary to all Area Offices. The policy direction was clearly stated: "Recent shifts in the economy have stimulated considerable interest by tribal leaders in the development of their woodland resources. This is primarily the result of opportunities which now exist for the sustained harvest of marketable forest products, particularly those associated with the Pinyon-Juniper woodlands of the west. Management of this resource is best provided Timber prices in the Southwest have greatly increased since this testimony was given. Nonetheless, the Pueblo's point about the need to focus on their predominant resource remains valid. under the direction of the Bureau's Forestry Program, as the dominant woodland values are relative to the culturing and utilization of true tree species, and as co-management principles, as well as multiple-use practices, are currently stressed nationwide by the program. You shall therefore adjust your programs accordingly and insure that all future resource management plans and plan revisions include the adaition of woodland management responsibilities to the Forestry Program. Arrangements should be made to provide assistance and oversight on these wooded reservations to meet the needs of the Indian owners." Several key points stand out in this directive. First, there was a recognition of tribal interest in woodland management. Second, woodland man- agement was moved under the direction of the for- estry program where the value of trees was more likely to be recognized and the resource managed in a multi-disiplinary manner. Third, the resource was to be actively managed and considered in for- est planning activities, rather than ignored as had often happened in the past. Finally, the focus of this planning and management was to meet the needs of the Indian owners. Recognizing that tribal goals and objectives for their woodland resources vary across the country, the Assistant Secretary provided latitude to respond to tribal direction in the management and oversight of woodland re- sources. This directive now guides woodland man- agement on Indian reservations throughout the United States. A few years later (1990), Public Law 101-630, the National Indian Forest Resources Management Act, became law. This legislation further cemented woodland's status by mandating proper planning and management of all Indian forest land. "Indian Forest Land" was specifically defined in the law to include both woodlands and timberlands. WOODLAND FUNDING AND RECOGNITION Congress appropriated the first woodland pro- gram dollars ($500,000) to the BIA in fiscal year 1990, thanks in large part to continued lobbying ef- forts by the Intertribal Timber Council and individ- ual Indian people. It is interesting to note that in 1990 the Director of the BLM signed an initiative called "Forests: Our 148 Growing Legacy/' giving woodland under their ju- risdiction the same status and policy guidance as other forest land; and in 1992, Region 3 of the For- est Service issued their Pihon-Juniper Initiative, also bringing woodland management into the limelight. Larry Henson (1993), the Southwest Regional Forester at the time, later publicly stated that the Forest Service "must improve pinon- juniper woodland management in the Southwest while recognizing the inherent values, including cultural values, of the pihon-juniper ecosystem." It is encouraging that the major land managing agencies are recognizing woodland's importance and are moving toward proper management of the same. It is also interesting to note that Mr. Henson pointed out the cultural connection inherent with woodlands. MANAGING WOODLANDS TO MEET TRIBAL NEEDS AND DESIRES Woodland ecosystems, given their long asso- ciation with people, need to be managed consider- ing the human component rather than considering man as an outside force. Archaeologists Cartledge and Propper (1993) state that in order to under- stand pinon-juniper ecosystems today, it is crucial to understand what those ecosystems were like in the past and what forces, including human beings, affected them. They continue, "Pinon-juniper woodlands did not exist in a pristine unaltered condition prior to the arrival of Europeans. Human beings have made extensive, and in some cases in- tensive, use of pinon-juniper woodland resources for thousands of years." Given that use, the key is to manage the woodland ecosystem in an integrated manner so that the resource is sustained even while it provides for the people dependent on that ecosystem. Even though the BIA's constituency (tribes and tribal members) is much more well defined than the Forest Service or the BLM constituencies, tribes also respond differently with requests for woodland funding and assistance. For example, the White Mountain Apaches have used woodland funding to open a wood yard on their reservation, but are reluctant to cut green pifion or juniper. Instead, dry Gambel oak (Quercus gambelli) currently makes up most of the fuel wood volume. On the other hand, their neighbors to the south, the San Carlos Apaches, recently harvested 840 cords of green alligator juniper (Juniperus deppeana) for sale through their tribal wood yard. Other Southwestern tribes are not interested in cutting any cordwood for resale to outside markets. The point is, tribes are given the latitude to customize woodland projects to fit local needs, desires, and traditions. Nonconsumptive uses of woodland areas are just as important to tribes as consumptive uses. Nonconsumptive uses include areas set aside for religious activities such as ceremonial dances (Utes) or planting prayer feathers (Zunis), or woodland areas set aside for wildlife habitat (Apaches) or scenic beauty (Navajos). The widespread practice of gathering pinon nuts — nonconsumptive, at least as far as the tree is concerned — also fits into this category. The vast majority of Southwestern tribes depend on pinon nut harvesting for traditional food or economic purposes. Although tribal mem- bers from many tribes do not sell pinon nuts as a matter of principle, Tanner and Grieser (1993) esti- mate that over 90% of the commercial pinon nut crop is harvested by Native Americans — a crop es- timated by Evans (1988) to be 1-2 million pounds of nuts per year. Due to cultural and tribal differences, wood- land project proposals requiring BIA funding must contain a signed tribal resolution to ensure tribal support and involvement. Proposals springing from the tribes have a much greater chance of suc- cess and allow diversity in management ap- proaches. Value-added processes or those that provide tribal employment are also looked at favorably when determining which projects should receive funding. The Zuni Furniture Enterprise is a perfect example of a tribal business that utilizes woodland resources and produces beautiful, value-added products while providing tribal income and em- ployment. The considerable artistic talents of the Zunis play a major part in making the furniture en- terprise a success. Four examples of diverse tribal approaches to woodland management and use from within the jurisdiction of the BIA's Phoenix Area Office follow. Hopi Reservation - Arizona Integrated Woodland Management Plan The Bureau of Indian Affairs is currently working with the Hopi Tribe to develop a compre- hensive, integrated woodland management plan for the tribe's 197,028 acres of pihon-juniper woodland. One of the first steps in the planning process involved a two-page questionnaire de- signed to poll tribal members about their priorities for tribal woodland areas. The questionnaire was 149 hand-carried to all of the Hopi villages by tribal members to encourage maximum participation. Re- sponses from 226 households were gathered. Tribal members were asked about their personal uses of the woodland areas, and what they thought the Tribe's management objectives for different wood- land areas should be. Respondents indicated that the top three management priorities for woodland areas should be for fuelwood production, cul- tural/religious needs, and water production. There is a strong desire to protect the woodland resource, but an equally strong desire not to infringe on tribal members' rights to utilize the woodland re- sources for ceremonial and subsistence purposes. "Religion cannot be regulated" was often heard during scoping sessions. The management plan will try to synthesize necessary human use re- quirements into an integrated management scheme that fully considers wildlife, watershed and soil protection, aesthetics, recreation, and range concerns. Uintah and Ouray (U&O) Ute Reservation - Utah Woodland Demonstration Blocks Six 2.5-acre woodland demonstration blocks were set up this year within a Pinus edulis/ Artemisia tridentata (PIED/ARTR) habitat type on the U&O Ute Reservation. Tree species present in the blocks are Colorado pinon (Pinus edulis) and Utah juniper (Juniperus osteosperma). The blocks demonstrate dif- ferent silvicultural prescriptions such as shelter- wood, group selection, a pinon nut orchard and a control block. Two other blocks show individual tree selection prescriptions with residual basal ar- eas of 40 and 70 square feet per acre, respectively. Fuelwood and fenceposts generated by the project will be transported to the tribal wood yard, proc- essed, and sold. Slash will be lopped and scattered to control erosion and to provide favorably mi- crosite conditions for grass, forbs and seedling re- generation. The entire area is to be fenced and signed after harvest. Plans also call for two additional blocks to be surveyed outside the fence. One of the blocks outside the fence will be a second control block to monitor any changes the fence itself causes, and one block will be installed in an adjacent area that burned in 1989. Installation of range and wildlife transects is included in the monitoring efforts. The blocks will provide a visual demonstration to the Tribal Council and tribal members and allow for lo- cal monitoring of changes resulting from silvicul- tural manipulation of the stands. Local site data gathered on this project will be useful in future decisions guiding proper management of the woodland resource. Washoe Pine Nut Allotments - Nevada Pinon Nut Production The Pine Nut Allotments in western Nevada's Pine Nut Mountains were reserved specifically for the Washoe people under the authority of the 1887 Dawes Act. Currently, there are over 66,000 acres in individual 160-acre parcels that belong to specific Indian owners rather than to a tribal entity. The allotments contain stands of singleleaf pinon (Pinus monophylla) with a smaller component of Utah ju- niper {Juniperus osteosperma). Problems associated with the stands include trespass cutting of live trees, a minor outbreak of Ips bark beetles (Ips con- fusus) and some dwarf mistletoe (Aeuthobium dwari- catum). This area provides a perfect opportunity to use good silviculture to protect the health and vigor of a resource that is culturally important. Treatments are planned which are designed to in- crease nut production and the vigor of the residual stand while reducing the incidence of dwarf mis- tletoe. Harvesting must be timed in late fall or early winter so as not to increase the potential for a large Ips outbreak. Allottees provide the specific goals and objectives or desired future condition for their allotment, while resource professionals help them achieve those goals through sound forest practices. Colorado River Indian Reservation - Arizona and California Woodland Inventory to Guarantee Sustainability of Cultural Resource The Colorado River Indian Reservation encom- passes 269,921 acres in Arizona and California bi- sected by the Colorado River. Tribal membership includes 3,057 people. This reservation is outside the range of pinon-juniper, but contains honey mesquite (Prosopis glandulosa) woodland. Many of the original stands of mesquite on the Reservation became depleted due to over- harvesting, changes in groundwater and flooding, and conversion of natural vegetation to farmland. The Colorado River Indian Tribe became concerned that the resource might not be able to meet traditional cultural needs and, therefore, placed a six month moratorium on mesquite wood cutting and land conversion. The cultural need, in this 150 particular case, was for the use of mesquite logs and arrowweed (Tessaria sericea) in funeral pyres for traditional Mohave cremations. Log lengths, diameters, and volume requirements of the mesquite are specific depending on whether the deceased member was an adult or a child. Given guidelines laid out by the Tribe, the University of Arizona's Office of Arid Lands Studies conducted a woodland inventory and issued a final report (Nabhan et al. 1985). Researchers had to match tribal death rates by size of the individual with woodland volume, growth data, and size of the trees. Although the subject is not a pleasant one to contemplate, the study was essential to Mohave tribal members who justifiably maintain that they must be guaranteed sufficient, sustainable quantities of mesquite and arrowweed for their funeral rites. As a result of the study, two large areas containing sufficient mesquite to meet the Mohave's future needs have been reserved along the banks of the Colorado River. CONCLUSION Hopi's integrated woodland management plan, the Ute's woodland study blocks, stand improve- ments designed to increase the Washoe's pinon nut crop, and the study to ensure a sustainable supply of mesquite for the Mohaves provide examples of managing woodlands to meet tribal needs and de- sires. The BIA's woodland program sprang directly from tribal concern and involvement. Tribes re- quested that woodlands be fully considered and properly managed rather than ignored or eradi- cated as often happened in the past. Tribes prefer a holistic approach to woodland management, taking into account the human element and cultural and spiritual requirements. However, tribes are also unique, with their own histories, languages, cul- tures, and religious beliefs. The BIA does not try to standardize woodland programs. Instead, diversity and unique tribal approaches within the guidelines of sound forestry practices are encouraged. ACKNOWLEDGEMENTS The author wishes to acknowledge the people of the Pueblo of Zuni. Living and working for four years with the Tribe was instrumental in shaping my own appreciation for the pifion-juniper wood- lands of the Southwest. Special thanks to the folks who spent time giving me personal insights into their tribal culture and beliefs. Appreciation is also extended to Lucille Wata- homigie of Hualapai who sent me two of her beau- tifully done books: Ethnobotany of the Hualapai and Ko (pihon). LITERATURE CITED Ackerly, Neal. 1991. Ethnobotany. In: Proceedings- 1991 Pinon Conference. 1991 April 23; Santa Fe, NM. New Mexico State University, College of Agriculture and Home Economics. Las Cruces, NM: 59-62. Arnold, Joseph E; Jameson, Donald A.; Reid, Elbert H. 1964. The pinyon-juniper type of Arizona: Effects of grazing, fire, and tree control. Prod. Res. Report No. 84. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 28 p. Aro, Richard S. 1975. Pinyon-juniper woodland manipu- lation with mechanical methods. In: The pinyon- juniper ecosystem: A symposium; 1975 May; Logan, UT: Utah State University, College of Natural Re- sources. Logan, UT: 67-73. Born, J. David; Tymcio, Ronald P; Casey, Osborne E. 1992. Nevada forest resources. Res. Bull. INT-76. Og- den, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station. 64 p. Bureau of Indian Affairs. 1988. Native American wood- land resources: A national overview, Assessing the re- source potential and management needs. U.S. Department of Interior, Bureau of Indian Affairs, Branch of Forest Resources Planning. Portland, OR: 139 p. Cartledge, Thomas R.; Propper, Judith G. 1993. Pinon- juniper ecosystems through time: Information and insights from the past. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings — Managing pinon-juniper ecosystems for sustainability and social needs; 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. De- partment of Agriculture, Forest Service, Rocky Mt. Forest and Range Experiment Station: 63-71. Conner, Roger C; Born, J. David; Green, Alan W; O'Brien, Renee A. 1990. Forest resources of Arizona. Res. Bull. INT-69. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station. 92 p. Dalen, Raymond S.; Snyder, William R. 1987. Economic and social aspects of pinyon-juniper treatment — Then and now. In: Everett, R.L., compiler. Proceed- ings— Pinyon-juniper conference; 1986 January 13- 16; Reno, NV. Gen. Tech, Rep. INT-215. Ogden, UT: U.S. Department of Agriculture, Forest Service, In- termountain Research Station: 343-350. DeBloois, Evan I.; Green, Dee F; Wylie, Henry G. 1975. A test of the impact of pinyon-juniper chaining on ar- chaeological sites. In: The pinyon-juniper ecosystem: 151 A symposium. 1975 May; Logan, UT: Utah State Uni- versity, College of Natural Resources. Logan, UT: 153- 160. Despain, Del W. 1987. History and results of prescribed burning of pinyon-juniper woodland on the Huala- pai Indian reservation in Arizona. In: Everett, R.L., compiler. Proceedings — Pinyon-juniper conference; 1986 January 13-16; Reno, NV. Gen. Tech, Rep. INT- 215. Ogden, UT: U.S. Department of Agriculture, For- est Service, Intermountain Research Station: 145-151. Evans, Raymond A. 1988. Management of pinon-juniper woodlands. Gen. Tech. Rep. INT-249. Ogden, UT: U.S. Department of Agriculture, Forest Service, Inter- mountain Research Station: 34 p. Henson, Larry. 1993. The Forest Service's pinon-juniper initiative for the Southwest. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceed- ings— Managing pinon-juniper ecosystems for sus- tainability and social needs; 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 23. Hurst, William D. 1977. Managing pinyon-juniper for multiple benefits. In: Ecology, Uses, and Management of Pinyon-juniper Woodlands. 1977 March 24-25; Al- buquerque, NM. Gen. Tech. Rep. RM-39. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Sta- tion: 45-47. Koyiyumptewa, Bruce K. 1993. Spiritual values of the pi- non-juniper woodland: A Hopi perspective. In: Al- don, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings — Managing pinon-juniper ecosystems for sustainability and social needs. 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 19-20. Lanner, Ronald M. 1981. The pinon pine: A natural and cultural history. University of Nevada Press: Reno, NV: 208 p. Lanner, Ronald M. 1993. What kind of woodland does the future hold? In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings — Managing pinon-juniper ecosystems for sustainability and social needs. 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. De- partment of Agriculture, Forest Service, Rocky Moun- tain Forest and Range Experiment Station: 14-18. Little, Elbert L. Jr. 1991. Pinon (Pinus edulis): An over- view. In: Proceedings-1991 Pinon Conference. 1991 April 23; Santa Fe, NM. New Mexico State University, College of Agriculture and Home Economics. Las Cruces, NM: 72-76. Little, Elbert L. Jr. 1993. Managing southwestern pinon- juniper woodlands: The past half century and the fu- ture. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings — Managing pinon-juniper ecosystems for sustainability and social needs. 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 105-107. McNichols, Robert R. 1987. Management strategies in pinyon-juniper on the Hualapai Indian Reservation. In: Everett, R.L., compiler. Proceedings — Pinyon- juniper conference; 1986 January 13-16; Reno, NV. Gen. Tech, Rep. INT-215. Ogden, UT: U.S. Depart- ment of Agriculture, Forest Service, Intermountain Research Station: 161-164. Miller, Ronald K.; Albert, Steven K. 1993. Zuni cultural relationships to pinon-juniper woodlands. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings — Managing pinon-juniper ecosystems for sustainability and social needs; 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Sta- tion: 74-78. Nabhan, Gary P.; Warren, Peter L.; Parton, Michael. 1986. Mesquite resources of the Colorado River Indian Tribes Reservation. University of Arizona, College of Agriculture, Office of Arid Lands Stud- ies. Tucson, AZ. Tanner, Ellis; Grieser, Don. 1993. Four generations trad- ing pinon nuts with Native Americans: Changes needed for future prosperity. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceed- ings— Managing pinon-juniper ecosystems for sus- tainability and social needs; 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 29- 33. Watahomigie, Lucille J.; Powskey, Malinda; Bender, Jorigine. 1982. Ethnobotany of the Hualapai. Huala- pai Bilingual Program, Peach Springs School District No. 8. Peach Springs, AZ. Watahomigie, Lucille J.; Watahomigie, Philbert Sr.; Powskey, Malinda; Bender, Jorigine; Uqualla, Josie. 1983. Ko. Hualapai Bilingual Program, Peach Springs School District No. 8. Peach Springs, AZ. 152 The Capulin Pinon-Juniper Ecosystem Management Project The Archaeological and Ecological Components John C. Phillips1 and Martha D. Yates Ph.D.2 Abstract. — The 1993 Capulin Pinon-Juniper Ecosystem Management Proj- ect is the culmination of a truly interdisciplinary planning process involving all resource disciplines and local community interests. Site selection incor- porated range condition and management, vegetation composition and structure, soil and hydrologic conditions, and accessibility by the public. The stabilization of significant prehistoric Gallina archaeological sites was incorporated into the management prescription. The primary objectives were to improve watershed condition, improve the biological and structural diversity of the Pinon-Juniper Woodland, provide a sustainable nut crop and fuelwood for the local community, and to stabilize Heritage Resource sites. The Pinon-Juniper ecosystems of the Southwest are now, and have been in the past a major source of agricultural land, food crops, fuelwood, forage for livestock, and building materials. This unique blend of natural resources and climate has favored human settlement. Many of the dwellings con- structed by Native Americans prior to European influences were located in these ecosystems; pres- ent-day villages and pueblos in Northern New Mexico often occupy the same areas. As a result some, if not most, of these areas have a long history of use spanning thousands of years (Cartledge and Proper 1993). Historic sites include homesteads, agricultural fields, and mine sites. Prehistoric sites can include pit house, lithic and sherd scatters, and agricultural garden systems. The demands for water, forage, and other resources continued into this century with an emphasis on livestock grazing and fuelwood production. At the turn of the century Pinion- Juniper was heavily harvested in the Jemez Mountains to provide fuelwood for the railroads used to transport timber. In the 1950's and 1960's, hundreds of thousands of acres of Pinon-Juniper were converted to grassland via chaining or dozer pushes. The grazing pressure in the first half of this century (Burkhardt and Tisdale 1976), combined with subtle climate shifts and drought (Betancourt et. al. 1993), and the lack of natural fires (Miller and Wigand 1994) have altered the conditions of these ecosystems. Many of the landforms that now support Pinon-Juniper ecosystems have higher than normal tree densities, closed canopies, and oppressed understories. The soil erosion exceeds long term rates (Pitlick, Wilcox and Allen 1994). Today, accelerated erosion over large areas exceeds rates that threaten site productivity (TES 1991). In some cases, well advanced gully and sheet erosion is causing increased mortality of long-lived species through exposure of root systems. The stability of archaeological sites associated with these ecosystems is often threatened. Today some of these areas continue to experience unauthorized fuelwood harvest, and uncontrolled livestock grazing which frequently exacerbates these existing conditions. The need to improve the management and conditions of these ecosystems is often obvious, but it is very complex. Managing these resources for sustainability, biological diversity, watershed in- 1 Zone Soil Scientist, Santa Fe National Forest, Coyote Ranger Station, Coyote, New Mexico. 2 District Archaeologis, Santa Fe National Forest, Coyote Ranger Station, Coyote, New Mexico. 153 tegrity, and wildlife habitat while preserving the values and heritage of Native Americans and oth- ers who use these resources is a very difficult task The Capulin Pinon-Juniper Ecosystem Man- agement Project is a significantly new approach to managing the Pinon-Juniper resources for the Coyote Ranger District on the Santa Fe National Forest. The project is unique in several ways. The planning process and management initiatives were developed using the concepts found within the Pinon-Juniper Initiative (USDA Forest Service Region 3), encompassing the needs of the public and traditional ways of life, while restoring sus- tainable ecosystem function. An ecological ap- proach was employed to develop the desired fu- ture conditions of the area. Biological, ecological, edaphic, and archaeological information from the project area were used. This information indicated the conditions of the stand prior to the turn of the century. The age class and structural diversity of the overstory, understory composition and expres- sion, soil productivity and conditions, and the past uses of the area were employed in order to deter- mine the conditions of the stand prior to European effluences. The desired future conditions were generated from this information and were utilized to develop management strategies for the stand. The Heritage Resources of the area were viewed as a component of the ecosystem. Instead of identify- ing and avoiding the Heritage Resources, their conditidns were incorporated into the planning. Management objectives were to provide fuelwood to the communities of Gallina, Coyote, and Regina, New Mexico while improving biological diversity, providing for improved soil stability, long term productivity, and overstory structural and species diversity. SITE SELECTION The potential natural vegetation, present vege- tation composition, soils, climate, accessibility by the public, fuelwood volume, livestock manage- ment, and site conditions were used when select- ing the project area. The Pinon-Juniper on the Santa Fe National Forest ranges in species domi- nance and understory composition along a climate gradient. At the cooler moister end of the climate gradient, at higher elevations, Pinon-Juniper is often co-dominant with Ponderosa Pine. Juniper Savannas, with little or no pinon in composition, occur on the warmer and drier areas at lower ele- vations. We sought to maximize the potential vege- tative response by locating the project in an area on the cooler, moister end of the climate gradient. The moister end of the climate gradient is more likely to have annual rainfall patterns and precipitation which will support a positive vegetative response and seedling establishment when compared to the more xeric conditions at the lower elevations. The Santa Fe National Forest contains Pinon- Juniper which is in relatively good condition, as well as areas which have been severely affected by past use and the subsequent erosion and/or higher than normal tree densities. Areas in good condition often have multiple-aged overstories with uneven distribution across the landscape. The canopy of the overstory often contains large masted Pinon and Juniper trees with canopy closures of less than 15 percent. The Pinon-Juniper stands in the best conditions have standing dead and fallen trees with an abundant herbaceous understory com- posed of native perennial grasses and forbs. In these conditions, it may only require the reintro- duction of fire to maintain these systems. On the other hand we have Pinon-Juniper ecosystems that are severely eroded. In these areas the soils have often been truncated by sheet or gully erosion that has resulted in increased mortality of the overstory due to exposed root systems and a loss of the her- baceous understory. A third condition exists: areas with good vegetative composition and site produc- tivity which are experiencing tree encroachment and accelerated soil erosion. These ecosystems have yet to cross the "threshold" of stability and sustainability where the system will quickly un- ravel. It was these systems "at risk" that we sought for the project area. One of the primary objectives of the project was to provide fuelwood to the residents of Coy- ote, Gallina, Regina, and Youngsville, New Mexico. This necessitated choosing an area of sufficient volume of cord wood for a public fuelwood har- vest, where public access could be controlled with existing fencing. It was also critical that the selected area had controlled livestock grazing with good compliance from the permittees. This was neces- sary if we were to achieve a favorable and sus- tained improvement in the herbaceous understory. The area selected had to contain a mix of per- ennial warm and cool season grasses and forbs which would respond favorably to the treatment and an overstory with a multiple age class distri- bution which had canopy closures sufficient to suppress the expression of the understory. We wanted a site which was experiencing accelerated erosion (gully and sheet/rill erosion). The erosion patterns had to be treatable and not sufficiently advanced to have totally truncated the soils 154 (having a total loss of the surface horizons), or gullies which had reached the top of the drainage basins or had down-cut to bedrock. In short, we sought to treat an area that was at risk of loosing its soil productivity to erosion, was sufficiently productive to have a positive response to the treatment, and contained all the necessary pieces (biotic and abiotic) necessary for establishing the desired future conditions. It was also critical that slash treatment provided for improved effec- tive ground cover, and a suitable micro- environment conducive for release and establish- ment of the herbaceous understory. PROJECT AREA The project location is in a Pinon-Juniper woodland adjacent to the village of Gallina, New Mexico. The area is approximately 7400 feet in elevation on a hill side slope with very shallow to moderately steep slopes. The area is located adja- cent to the Capulin Creek which is a perennial tributary to the Gallina River. The area is composed of a mature multi-aged Pinon-Juniper woodland with an understory of Gamble Oak, Big Sagebrush, and native perennial grasses. The Terrestrial Ecosys- tem Survey (TES) for the Santa Fe National Forest describes the soils of the area as Typic Haplustalfs, fine loamy, mixed, mesic; moderately deep and deep sandy loams. These soils support a potential natural vegetation of Pinus edulis, Juniperus mono- sperma, Quercus gambelii, and Artemisia tridentata spp. tridentata (TES LSC4 + 1). The area is on the cooler moister end of the climate gradient that supports Pinon-Juniper ecosystems. Understory species present in composition included Agropyron smithii, Muhlenburgia montana, Koeleria macrantha, Sporobolus contractus, Schizachyrium scoparius, Bouteloua gracilis, Oryzopsis hymenoides, Sitanion hystrix, and Blephar- oneuron tricholpis. Trace amounts of Bromus tectorum are present on the more disturbed sites within the planning area. Big sagebrush is present throughout and varies in expression depending on the canopy closure of the overstory. The overstory is composed predominantly of Pinus edulis and Juniperus mono- sperma with lesser components of Juniperus scopulo- rum and Juniperus osteosperma. Trace amounts of Pinus ponderosa, and Pseudotsuga menziesii are pres- ent in topographic positions on the land form. The geology of the area is composed of the upper shale member of the Chinle Formation and localized terrace and pediment deposits. The dominant erosion processes are sheet and rill erosion which have resulted in pedistalled plants and debris dams. Well developed shallow gully networks were often associated with archaeological sites. Many of the gullies were threatening the stability of pithouses and other heritage resources. Many had reached the top of the basins in which they were migrating. Secondary headcut development was observed in existing gully networks. The overstory varied in canopy closure from between 60 and 85 percent. Small areas of even-aged mature Pinon-Juniper with canopy closures of 50 to 70 percent were present. The understory expression was severely limited by the closed canopies or erosion processes. Herbaceous canopy cover ranged from 4 to 16 percent. Bare soil was less than 25 percent within the closed canopy areas due to the heavy needle cast; however, where the canopy was less dense due to unauthorized fuelwood harvest, the presence of archaeological sites, and/or roading, bare soil was as high as 60 percent. In the past, areas with high densities of heri- tage resource sites were avoided for fuelwood sales. On the Coyote Ranger District, high densities of prehistoric sites are very common and are almost always present within the Pinon-Juniper ecosys- tems. Although for this project we did not seek an area of high site density, we did recognize that wherever we planned treatments in the Pinon- Juniper ecosystem, Heritage Resources will be present. By bringing archaeological sites into the management prescription, we were able to treat the area on a landform basis, and provide for long term protection of the Heritage Resource sites. THE HERITAGE RESOURCE COMPONENT The Capulin Pinon-Juniper ecosystem man- agement project is located in a region dense with Gallina culture settlements dating from 1059 A.D. to 1300 A.D. Much of what has been variously called the Gallina Phase (Hibben 1938), Largo Phase (Mera 1935), or Largo-Gallina Phase culture is located in the Coyote Ranger District of the Santa Fe National Forest. The Rosa Phase (ca. 700 to 850 A.D.) of the Gobernador Canyon and Navajo Res- ervoir area (San Juan River) appears to be the most likely ancestor to the Gallina settlements, although the temporal hiatus between 850 A.D. and the beginning of the Gallina phase (1059 A.D.) has not yet been explained. The settlements of the Gallina make up the majority of the Heritage Resource sites located within the Capulin Pinon-Juniper Ecosystem Proj- ect Area. Six centuries have passed since these 155 people abandoned their villages, yet their pithouses remain, as well as significant amounts of cultural materials. Ceramics, lithics, remnants of wooden stockades and stone towers, storage bins, hatch covers, clay elbow pipes, pointed -bottom pots (unique to the southwest), tri-notched axes, worked antler, and comb arrow polishers are some of the structures and materials which remain. The Heritage Resource survey which was completed in July of 1993 located twenty-one ar- chaeological sites within the 160-acre project area (Yates 1993). Eight sites are Gallina pithouse clus- ters, twelve are sherd and lithic scatters, and one site included both a lithic scatter and the remnants of the USFS Capulin Ranger Station (ca. 1912-1932). Gallina pithouses are subterranean dwellings, measuring between 7.5 meters and 13.5 meters in diameter and approximately four meters deep. Two of the twelve pithouses in the area have a surface storage structure attached. The cultural materials located during the survey included ce- ramics (Gallina black/white and grey ware), lithics (projectile points; primary, secondary, tertiary flakes; cores; and BR/M flakes), hatch covers used for Gallina storage bins, flagstone, and sandstone and basalt grinding stones. Their agricultural systems could be quite well developed and are, culturally, very significant. Within the Coyote Ranger District, there are a number of Gallina agricultural features: terraced gardens, grid gardens, and reservoirs. In the Rincon Colorado area along the mesas of the Gallina River, each dwelling structure is associated with an area of terraced gardens. On Mesa Golondrina, high above the junction of the Chama and Gallina Rivers, a Gallina settlement called "Castles" includes pithouses, a pueblo-like structure, surface houses, and a possible tower. On the lower mesa of Golondrina, there are terraced gardens and an ancient reservoir (Douglas 1917 and Dick 1976). At Rattlesnake Ridge in the Cuba District of the Santa Fe National Forest (both Rattlesnake and Castles are listed on the National Register), an ancient Gallina Reservoir (at Bg 20/2) is estimated by Turney (1985) to have held more than 29,000 gallons of water, enough water to supply 1.75 gallons/day to 150 people for 45 days without replenishment (see also, Perret 1976). Yet, despite the significance of the Gallina cul- ture, only a small number of Gallina agricultural features have been studied. Few of their dwellings have been excavated. Of those which have been excavated, many were found to have been burned (Mackey and Green 1979). Charred remains of individuals were present within the structures. At Rattlesnake Ridge during a pithouse excavation, ten adult skeletons were found in a pile and each skull had pre-mortem fractures. Four of the skele- tons had projectile points embedded, and there was some evidence of cut marks in the bone. This kind of evidence suggests some form of internal or external strife. Much about the Gallina people remains a mystery, making the preservation of all Gallina Heritage Resource sites extremely important. The origin of the culture is not completely understood, nor is its demise. We do not know where they disappeared to or if they were, indeed, overrun by an incoming people. No one knows with what modern-day peoples they are culturally affiliated. None of the modern pueblos along the Rio Grande claim the Gallina as their ancestors. Nor do we know the specific ways in which an agricultural technology (probably brought from Mesa Verde or the San Juan River) was adapted to the higher elevations and differing climatic conditions of this region west of the Chama River and north of the San Pedro Mountains. The Coyote Ranger District has the immense task of managing and protecting the remnants of a culture which has not yet been adequately researched or understood. Yet, some of the earliest archaeological research on the Gallina culture was based on excavation and survey work conducted within what is now the Coyote District's Capulin Pinon-Juniper Ecosystem Management project. Much of the research for H. P Mera's article in American Antiquity called "Some Aspects of the Largo-Gallina Phase, Northern New Mexico" was compiled during the 1934-35 season when his team excavated a pithouse (LA 641) located in one of the three pithouse clusters within the Capulin project area (Mera 1938). The excavated pithouse with its subterranean hearth, deflector, ventilator shaft and storage bins dates to 1106 A.D. From the Mera excavation we know that over nine hundred years ago, ancient people established villages in the Pinon-Juniper woodland along the Rio Capulin and were utilizing its resources. For over 10,000 years, human beings have been part of the Pinon-Juniper story. These woodlands contain the highest densities of Heritage Resources found anywhere in the Southwest (Cartledge and Propper 1993). Coprolite analysis has shown that the pirion nut was second only to corn as a food source (Aasen 1984). The Pinon-Juniper woodland contains a tremendous variety of foods: yucca, oak, rice grass, prickly pear, and juniper are some of these. Within the pinon-juniper woodlands of Black Mesa (Arizona), thirty-four useful plant 156 species occur: twelve are food sources, four are medicinal, and eleven are used as raw materials (from Cartledge and Propper 1993, based on Gu- merman 1984). Within the Capulin ecosystem management project, thirteen of the Heritage Resource sites are tool maintenance and reduction areas, containing thousands of flaked artifacts, the remnants of the stone toolmaking process which in the upper Chama River region and eastern Jemez Mountains dates to the Paleo-Indian period (pre-5000 B.C.). During the project Heritage Resource survey, an Archaic component to the Capulin Pihon-Juniper woodland was documented when an En Medio- Parallel projectile point (1000 B.C. to 200 A.D.) was located (Yates 1993). Today, the southeastern corner of the Capulin Ecosystem Management Project area is only a few hundred yards from the Hispanic village of Gallina and its elementary and high schools. For decades the present-day local communities have been using the area for fuelwood cutting and pifton nut gath- ering, much as the Gallina people almost ten cen- turies ago, utilized and eventually, possibly, de- pleted its resources. By considering the ancient human dimension, we can understand how these woodlands have come to be as they are. They are "not pristine and unmodified" (see Cartledge and Propper 1993). Archaeological research which includes paleoenvironmental, faunal and floral research can tell us much about the effects past cultures have had on the health of the Pihon- Juniper woodland. This, in turn, can help us make choices when managing the Pihon-Juniper wood- land for the benefit of present-day communities. The high use of the Pihon-Juniper woodland in the distant past and its usefulness now makes the management of these areas a challenging, if not an immensely complex task The 160 acres of Pinon- Juniper woodland along the Rio Capulin selected for this ecosystem project exhibits the elements which make Pihon-Juniper management difficult if not daunting (high density heritage site distribu- tion in an area which supplies local communities fuelwood and grazing land). Continuous use over many thousands of years has promoted erosional patterns which threaten both the heritage re- sources and the production of important natural resources upon which local people depend. At the onset of this project, soils in the area were disap- pearing, pihon root structures had been exposed, in some areas grass production was non-existent, and gully washes one and a half meters deep ex- isted within eight meters of Gallina pithouse set- tlements. The routine avoidance of the twenty-one heri- tage resource sites within the Capulin project area would not have helped preserve them. In fact, without the mitigating treatments used in the Coyote ecosystem Pihon-Juniper project, two of the pithouses and several of the lithic scatters in the area eventually would have been lost. Roads which traverse pithouses, used for years for pihon cutting, had caused continuous damage. As part of the Capulin project, the use of all roads through the area has now been controlled. The Heritage Re- source sites have been seeded, grass production has been improved, and gully washes within sites have been stabilized. The primary objectives of the Capulin ecosys- tem management project were to improve the size and age class diversity of the stand, improve soil and watershed conditions, and to protect and stabilize Heritage Resource sites from further deg- radation due to erosion. A fuelwood sale was ad- ministered during which the public was not al- lowed to cut within the heritage sites. However, the New Mexico State Historic Preservation Officer approved the stabilization plan with a provision which allowed Forest Service personnel to cut within the sites and spread the slash. The stipulations detailed in the Heritage Re- source clearance report were stringent and, in some cases, unique to different sites (Yates 1993). The slash cut from trees within the sites would be distributed in such a way as to stop or, at least, slow sheet and gully erosion and to encourage seedling development. All slash would be lifted and not dragged into place. Cordwood would be carried outside the boundaries of the sites. The sites would be cordoned off with bright yellow flagging; no vehicular traffic was allowed, public or Forest Service. The archaeological sites would be broad- cast seeded by hand. RESULTS The desired future conditions of the areas were to move the vegetative structure and composition back to a more open uneven-aged Pihon-Juniper Woodland favoring the masted healthy crowned nut producing trees. The spacing on the leave trees was determined by intensive field reconnaissance. Initially the spacing was to be between 80 and 100 feet on the mature masted trees with smaller di- ameter trees interspersed between. During field reconnaissance it was determined that the favored leave trees (the healthiest and largest as evidenced by their size and canopy condition) were on a 157 spacing of between 40 and 60 feet. Intermediate size trees greater than four inches in diameter were marked as leave trees to provide for an uneven- aged stand. The vegetative treatments were performed via a tightly administered public fuelwood harvest. The area was cruised to establish "woodlots" with a volume of two cords. Fifty (two cord) fuelwood permits were offered to local residents on a first come, first serve basis. Woodlots were allocated by a lottery. Permit requirements involved felling trees except for those marked as leave trees, utilization of coarse woody material down to a two inch diame- ter, lopping slash to a height of eighteen inches, and scattering slash to achieve an even distribu- tion. Treatment of the Heritage Sites was per- formed by all of the District resource areas through a District Work day. Soil protection was achieved by prohibiting fuelwood harvest when soils were wet (soil moisture greater than 25 percent of field ca- pacity), locating vehicle access routes on the con- tour, and restricting vehicle access to slopes less than twenty percent. The project was implemented in September, 1993. Monitoring the changes in ground cover (vegetation, bare soil, litter) and vegetative re- sponses was performed by a stratified random sampling of the ground cover components by pace transect, by permanent photo points with a known azimuth, and by qualitative ocular estimates. Ap- proximately 25 acres were treated including 6 to 7 acres of Heritage Resource sites. The immediate result was a decrease in bare soil from 21 percent to 11 percent for those areas with canopy closures greater than 60 percent prior to treatment. In the more open canopy areas bare soil was reduced from approximately 60 percent to less than 30 percent. Overstory canopy cover was reduced to less than 15 percent. Denned gullies within the treated area were stabilized using brush and debris dams and scattered slash. Sediment catchment was observed behind debris dams and behind slash. Several days after treatment, a rainfall of light intensity and moderate duration occurred, fol- lowed by several days of warm weather. Visual observations of the existing vegetation and mois- ture content of the soils were made several days after the precipitation. Individual plants protected by the slash snowed a growth response, where as plants not protected by slash showed no observ- able response. In addition, soil moisture was higher within and under the slash when compared to the moisture content of the soils not protected. The following summer (1994), during seed set of the cool season grasses, those plants protected by slash were physiologically superior to plants not protected by slash. Vegetation growth was more vigorous, plants were taller with more leaf area, and had larger and more numerous seed heads. The canopies of the plants protected by slash were greater than those not protected. Seedling estab- lishment was observed occasionally under slash, but more frequently within the sediment catch- ment behind debris dams. Approximately one third of the area treated by the public had slash densities restrictive to plant growth or seedling establishment. Although this condition restricted plant growth and establish- ment, it did provide for improved ground cover and soil stability. Public compliance with the lop and scatter criteria was good. No leave trees were taken and the use of smaller diameter trees was fair. Lopping was performed most of the time; while scattering of the slash was performed less completely. Comments by local community mem- bers indicated a more relaxed atmosphere, know- ing they had their own woodlot to harvest their fuelwood from. CONCLUSIONS The Capulin Pifion-Juniper Ecosystem Man- agement Project was able to be implemented be- cause its primary objective was the preservation of all resources: the Heritage Resources left behind by ancient people and the woodland resources upon which present-day local communities depend. It involved all the resource areas of the District — timber, range, heritage resources, soils, recreation, and fire — as well as local community interests. All shops within the Coyote Ranger District, as well as the Elder Americans, participated in this Heritage Resource stabilization and soil and watershed improvement project, and all resource areas bene- fited from the project. The goals of the different disciplines were the same: improve the quality of soils and herbaceous understory, retard erosion, and improve the age class distribution of the Pi- non-Juniper stand. These methods and conditions help slow erosional processes that threaten Heri- tage Resource sites. This was the first time the New Mexico Stage Historic Preservation Officer (SHPO) had ap- proved such a plan. It was the first time on the Santa Fe National Forest that fuelwood had been cut from within the boundaries of archaeological sites and site stabilization had been incorporated into a fuelwood sale. To our knowledge, it was the first time in the Southwest Region (USFS) a plan of 158 this sort — in which all District resource areas worked together to preserve Heritage Resource sites — had been implemented. Consideration of the human dimension was one of the primary goals of this ecosystem man- agement project, both the ancient human dimen- sions (archaeological site stabilization) and the modern-day one, providing fuelwood sales while ensuring sustainable ecosystem function. During this project, Heritage Resources and Watershed Management moved from their old function of project support to a new one, as full partners in ecosystem management. LITERATURE CITED Aasen, D.K. 1984. Pollen, Macrofossil and Charcoal Analyses of Basketmaker Corolites from Turkey Pen Ruin, Cedar Mesa, Utah. Unpublished Master's The- sis, Department of Anthropology, Washington State University. Pullman, WA. Burkhardt, J.W; Tisdale, E.W. 1976. Causes of juniper invasion in southwestern Idaho. Ecology. 76: 472-484 Betancourt, Julio L.; Pierson, Elizabeth A.; Rylander, Kate Aasen; Fairchild-Parks, James A.; and Dean, Jeffrey S. 1993. Paper presented at the Managing Pinon-Juniper Ecosystems for Sustainability and Social Needs Sym- posium. Santa Fe New Mexico. April 26-30, 1993. Gen. Tech. Rep. Fort Collins. CO. US Department of Agri- culture, Forest Service, Rocky Mountain Range and Forest Experiment Station. RM-236. Cartledge, Thomas R. 1988. Gallina Culture Develop- ments in North Central New Mexico. National Regis- ter of Historic Places Multiple Property Documenta- tion Form. USDA, NPS. Cartledge, Thomas R.; Propper, Judith G. 1993. Pinon- Juniper Ecosystems Through Time: Information and Insights from the Past. Paper presented at the Manag- ing Pinyon-Juniper Ecosystems for Sustainability and Social Needs Symposium. Santa Fe, N.M. April 26-30, 1993. Dick, Herbert. W. 1976. Archeological Excavations in the Llaves Area Santa Fe National Forest, New Mexico, 1972-1974. Ms. on file, Southwestern Region, Albu- querque, N.M. Douglas, William Boone. 1917. Land of the Small House People. El Palacio. 4:3-23. Santa Fe, N.M. Gumerman, George J. 1984. A View from Black Mesa: The Changing Face of Archaeology. University of Ari- zona Press. Tucson, AZ. Hibben, Frank C. 1938. The Gallina Phase. American Antiquity. 4 (2): 131-136 Mackey James; Green, R.C.. 1978. Largo-Gallina Towers: An Explanation. American Antiquity. 44 (1): 144-154 Mera, H.P 1938. Some Aspects of the Largo Cultural Phase, Northern New Mexico. American Antiquity. 3 (3): 236-241 Miller, Richard E; Wigand, Peter E. 1994. Holocene Changes in Semiarid Pinyon-Juniper Woodlands: Re- sponse to climate, fire, and human activities in the US Great Basin. Bioscience. Vol. 44 No. 7: 468-472 Perret, William. 1976. Report on BG20/2 Reservoir, Rat- tlesnake Ridge. Ms. on file, Ghost Ranch Conference Center, Abiquiu, N.M. Pitlick, John; Wilcox, Bradford; and Allen, Craig 1994. Observations of Runoff and Sediment Yield in a Pin- yon-Juniper Watershed in Bandelier National Monument, New Mexico. Unpublished manuscript. Turney, William F. 1985. Prehistoric Water Reservoirs — the Southwest. In the Collected Papers in Honor of Albert H. Schroeder. Southwest Culture History. 10: 43-57. Yates, Martha D. 1993. Capulin Fuelwood Sale and Soil and Watershed Management Project. Coyote Ranger District, Santa Fe National Forest. Ms. on file, Southwest Region, Albuquerque. Report no. 1993-10-41 159 Community Based Pinon-Juniper Woodland Resource Management Planning for the Nahat'a' Dziil Chapter Usha Little and Denver Hospodarsky1 Abstract. — The Navajo Nation's New Lands Nahat'a' Dziil chapter located near Sanders Arizona covers 350,000 acres of pihon-juniper woodland and is divided into 17 Range Management Units. The Office of Navajo and Hopi Indian Relocation currently manages these as functionally independent units emphasizing clustered residential development, livestock grazing, wildlife habitat improvement, and management of woodland resources for sustainable economic growth. Research in other counties suggests that community based resource management planning based on community forestry principles is a feasible option for the long-term sustainability of natural resources in the rural communities. The purpose of this on-going study is to evaluate the applicability of community forestry programs for the management of New Lands' woodland resources. The fundamental con- cepts of community forestry are to allow residents' needs and desires to define future conditions, and to incorporate community residents' rights and responsibilities of woodland resources stewardship into a management plan. The study has attempted to evaluate: (1) residents' recommendations regarding uses and allocations of woodland resources, (2) their willingness in assuming responsibility of developing community based plans and proj- ects, and (3) feasibility of achieving mutual goals between resource man- agers and residents through community forestry approach. Informational efforts were initiated to promote awareness of the community forestry con- cepts through presentations, small group discussions, and a small-scale community vegetation restoration project. Community input was satisfied through interviews of small planning groups of grazing permitees and resi- dents from each Range Unit. A series of meetings and communications were maintained to achieve coordination among residents, elected officials, community planners, and the resource managers. Some results and con- clusion are discussed. INTRODUCTION In recent years a community based approach to addressing forest management issues has gained momentum in the United States. Long established abroad, the community based approach has been slow to gain acceptance domestically, however. As a result, the forestry profession in this country has only begun to realize the benefits of forest resource decision making conducted at the community level 1 Native American Forestry Program, School of Forestry, Northern Arizona University, Flagstaff, AZ. with the goal to improve environmental and eco- nomic conditions for rural people. Recent expressions of concern for rural com- munity stability and rural resource development by federal land management agencies may be in- terpreted to suggest maturing appreciation for the community approach, at least among public land managers. Such recent developments seemingly contradict long standing statutory expressions of concern for resource dependent communities in 160 the United States beginning as early as 1944 under the Sustained Yield Act and continuing with the National Forest Management Act of 1976. Despite some recent developments in this country in sup- port of community based forest management, community forestry as it is formally practiced abroad has remained largely a conceptual exercise with few applied examples to draw from. As a result of the paucity of examples, any pro- posals to apply community forestry principles to North American land management must necessar- ily draw from conceptual origins within the foreign experience, while incorporating the unique traits of the domestic community under study. As the prac- tice of community forestry develops in the United States, each new attempt to implement the com- munity approach will require an initial assessment of its appropriateness for the specific population and resources in question. The need for empirically based community forestry principles are especially acute in our culturally diverse society whose indi- vidual communities reflect a multitude of needs, problems, and aspirations regarding the forest re- source. This paper reports on one attempt to de- termine the applicability of community forestry principles and methods, in this case pihon-juniper woodland management planning in New Lands communities of the Navajo Nation, Arizona. COMMUNITY FORESTRY Community forestry has been defined by Fisher et al.(1989) as a forest management practice that includes protection, utilization, and distribution of forest products, and the institutional and organ- izational arrangements by which the management prescriptions are carried out. In community for- estry the concern with institutional and organiza- tional arrangements involves a shift from the technical-driven classical top-down approach with its focus on forest productivity to a people- centered, bottom-up approach that emphasizes helping communities to manage the forest upon which traditional livelihood and culture may de- pend. Both the classical and community approaches to forest management share the same universal questions of production, protection, distribution, and management control. The community forestry approach, however, offers the opportunity to inte- grate local peoples' desires and knowledge, which has been largely constrained in the practice of tra- ditional forestry (Burch 1988). Contemporary defi- nitions of community forestry focus on the control and management of forest resources by local peo- ple who use forest resources for domestic purposes, as an integral part of traditional living (Fisher and Gilmour 1990). Community forestry projects ad- dress issues beyond lumber and fuelwood includ- ing the management of forest related products important to the community, such as fabric dyes, herbs, medicines, honey, fodder, nuts, and wildlife hunting. Community forestry incorporates an in- digenous system of management, which allows the incorporation of traditional values into the formu- lation and approval of resource management poli- cies that formally empower community groups. Under the community forestry model, wood- land management plans must consider technical forestry data in conjunction with knowledge of community social structure and cultural values as they relate to the woodland resource uses. The community forestry approach allows incorporation of local residents' knowledge, needs, and desires in defining future conditions, by incorporating resi- dents' rights and responsibilities for woodland re- sources stewardship into a management plan providing empowerment to local people. PLANNING AREA Historical Background of the New Lands In 1983, the Navajo Nation acquired 352,000 acres of Tribal Trust land, known as the New Lands, for the purpose of providing settlement to the families affected by the Navajo Hopi Settlement Act (Public Law 93-531) of 1974. The New Lands pro- vides a settlement option to any certified relocatee and their extended families who do not wish to be relocated off reservation and who do not qualify or wish to acquire a homesite lease in the main reser- vation, but wish to maintain the traditional Navajo lifestyle based on livestock grazing and large ex- tended families. The Office of the Navajo and Hopi Indian Relocation (ONHIR) has responsibility for managing the land and resources, while it contin- ues additional development and relocation activi- ties on the New Lands. The New Lands is located 40 miles east of Holbrook along Interstate 40 near Sanders in Northeast Arizona. The topography of the New Lands includes flat grassy plains, gently sloping benches and steep broken ground with the elevation varying from 5,500 to 6,900 feet. About 30 percent of the New Lands is considered wooded area covered with pihon-juniper woodlands. The entire New Lands is divided into 17 range 161 management units, 11 of which have already been developed with clustered homesites and grazing management cells for livestock. The clustered homesites are provided for the families with livestock grazing permits and permitees immediate relatives. Families without grazing permits are offered one acre lot homesite leases within the rural community sub-divisions of New Lands (ONHIR Plan Update 1990). All Navajo families relocated either into the range units or into the rural community sub-division have equal access to the woodland resources of the New Lands including firewood, pinon nuts, medicinal plants, scenic views, wildlife, and access to the religious sites. COMMUNITY FORESTRY ON THE NEW LANDS There are five reasons for selecting New Lands for community based woodland resource planning. First, New Lands currently does not have a sus- tainable management plan acceptable to commu- nity resource users. Second, New Lands' communities are recently established and commu- nity residents are fairly receptive to new ideas re- garding resource management that are outside the classical forest management model practiced else- where on the Navajo Nation. Third, the residents of the New Lands communities include Navajo people who use adjacent pihon-juniper woodlands as an integral part of traditional living. Fourth, many New Lands residents desire innovative, but culturally appropriate, options for employment and income opportunities from woodland re- sources, in addition to more traditional uses. And finally, New Lands people express substantial con- cerns for the welfare of the land, resources, and people after the ONHIR support discontinues in completion of the relocation activities. The above situation presented an opportunity to study the applicability of the community forestry approach on New Lands. DEFINING COMMUNITIES A primary step toward a community based woodland management planning approach is de- fining and delineating the community as the op- erational unit of social organization within which to undertake community forestry activities. There are three general approaches in defining commu- nity. One approach may be to define community as locality, a human settlement with fixed and bounded territory. This definition is a favorite of economists who analyze locational areas based on where people live and work. Another approach defines community as a local social system involv- ing interrelationship among people living in the same geographic area. This definition does not de- fine clear geographic areas and lacks definitions of content and quality of social interactions and inter- relationships. A third approach defines community as a type of relationship, especially a sense of shared identity without any geographic boundary, such as the environmental community (Lee et al. 1990). With the purposes of this study to assess the vi- ability of communities for the community based woodland resource management planning ap- proach, community is defined for our purposes here by integrating some elements from all the above three definitions. In this study community refers to a human settlement delineated by geo- graphical boundary and represented by locally or- ganized political and social groups, which maintain formal social interactions, while engaged in activi- ties to fulfill the needs and desires of the people holding common shared cultural values towards the land and pinon-juniper woodland resources. OBJECTIVES The purpose of this study is to determine the applicability of community forestry principles in the management of pinon-juniper woodlands in the New Lands communities. To evaluate the ap- plicability of the community forestry approach, three research elements were identified as funda- mental to community based woodland planning. These elements were 1) delineation of operational communities within the New Lands, to serve as the unit of analysis in resource planning research and plan implementation and analysis of existing insti- tutional structure and capability of communities to accommodate the community approach, 2) identifi- cation of local peoples' needs for , attitudes toward, and determination to manage woodland resources. This paper describes how each of these elements may be integrated into a community based ap- proach to pinon-juniper management planning for the New Lands. 162 SUPPORTING RESEARCH FOR COMMUNITY BASED PLANNING With the main objective of the study to deter- mine the applicability of the community forestry approach to pifion-juniper management planning for New Lands, three different kinds of research methods were used to meet planning information needs: 1. Formal interviews of community resi- dents designed to determine a) types, quantities and frequencies of woodland resources use by the residents, b) com- munity attitudes, opinions, and prefer- ences towards present and future woodland management planning goals and activities, and c) who or which community group should involved in woodland resource management plan- ning. 2. Direct observation of community activi- ties and interactions obtained during personal visits. Direct observation methods were used to determine a) willingness of the management agency and chapter officials to coordinate woodland resource related community activities, b) both residents and chapter officials' willingness to take active par- ticipation in vegetation restoration ac- tivities. 3. Secondary data including community demographics, existing livestock grazing activities, and woodland inventory in- formation pertinent to multiresource woodland resource management objec- tives. Successful application of the community for- estry approach requires adequate knowledge of the affected population and their relationship to woodland resources. Prior to conducting inter- views of New Lands residents, careful considera- tion was given to the population's relationship with pifion-juniper woodlands, according to vari- ous ethnohistories commencing both before and following the arrival of anglo-Europeans. In addi- tion, the sensitive nature of relocation activities was carefully considered in developing survey ques- tionnaires and in identifying effective methods for approaching residents for interviews. It is vital for the community forestry researcher or planner to understand and respect community social norms, religion, and traditional values before embarking upon planning for woodland resource issues. Besides knowing about the Navajo people, participating in community social activities, and in- teracting with the people as a friendly help instead of authoritative figure is essential. It is necessary to create a spirit of mutual trust and respect for the success of this study or any other successful com- munity forestry enterprise. The community based process of managing pinon-juniper woodlands de- scribed in this paper is built upon a profound ap- preciation for evaluating Navajo culture and traditional values, as these values were used in identifying pertinent community forestry variables and interpreting study results. RESULTS AND CONCLUSIONS Personal interviews of New Lands residents indicate that the woodland resource is subject to both traditional and non-traditional demands. Traditional demands condoned by most residents include harvesting fuelwood and postwood for the personal uses, collecting juniper seeds and leaves for ceremonies, pihon nuts, and medicinal plants. Non- traditional demands include residents' desires for economic development activities through retail- ing fuelwood, postwood, and other marketable woodland resources. When asked whether New Lands juniper should be harvested as fuelwood to create jobs and income for the New Lands resi- dents, 59 percent responded affirmatively and 39 percent said no. The survey data indicated a community based forestry approach is desired by the majority of the New Lands residents and elected officials. More than 50 percent of the interviewed residents ex- pressed willingness to work with a management agency in developing plans that incorporate their opinions into the planning process. Among inter- view respondents, 28 percent mentioned that final decisions about woodland resources should be made only by community people; while 22 percent said that Navajo Forestry Department should work with local planners, chapter officials, and commu- nity people in the planning process. The survey results indicated, woodland re- sources are important to the New Lands people in their everyday life and that both New Lands elected officials and the residents prefer residents' opinions and needs be honored in the decision making process. The majority of those interviewed responded favorably to a locally developed plan that empowers local people. When asked if they were willing to organize a community group to work on developing woodland management 163 planning, 60 percent of those community residents interviewed responded favorably, while 20 percent were not willing and 14 percent responded that they are too busy to organize such group. Community forestry in practice must be based on clear delineation of what and where the subject community is, in order that forestry program comprehensively address the resource needs and desires of community residents, while considering the abundance and sustainabiiity of woodlands within the community resource area. What practice requires is an operational defini- tion of community that insures the community forestry program will proceed efficiently to- wards the mutually agreed upon goals of com- munity residents and resource management agencies with stewardship responsibility for community resources. Four levels of existing social organization of- fer promise as the basis for delineating opera- tional communities in the New Lands. In descending order of organizational complexity these are: the Nahat'a' Dziil Chapter, Planning Board, range units, and grazing committees. The Nahat'a' Dziil Chapter and Planning Board are recognized as political organizations whose rep- resentatives are elected by Chapter residents. Chapter officials consist of a President, Vice President, and Secretary Treasurer. A monthly chapter meeting is held in which New Lands residents are encouraged to participate. The Planning Board includes two representatives from each range unit. Representatives are ap- proved the residents of the respective range units to represent range unit issues to Chapter officials. The Planning Board's operational or- ganization is unique to the New Lands; other Chapters of the Navajo Nation may not include such an organizational structure. Range units are characterized by groups of families living within the geographic boundary of prescribed range units. Grazing committees comprised of mem- bers from within individual range units, are rec- ognized as a traditional social organization of Navajo culture. Grazing committees are com- prised of families who have inherited livestock grazing rights to the land. There are various ad- vantages and disadvantages associated in rec- ognizing each of these social units as a basis for delineating operational communities within which to initiate community based woodland management planning approach. These advan- tages and disadvantages can be summarized as follows: Chapter Government as Community Basis The operational community organized around the Chapter can provide advantages such as the unity and strength displayed within the Chapter as whole, political authority for implementation of projects, monitoring, and widespread distribution of benefits. However, there are several disadvan- tages of using the Chapter level of organization as a basis for delineating community, such as resi- dents' mistrust of Chapter authority, since the Chapter may hold political motives as a result of its direct ties to Tribal government. Chapter officials change every four year and priorities given by Chapter officials in defining community needs may changed depending on the personalities of the elected officials. Chapter officials have many re- sponsibilities and may not find enough time to talk with people and subsequently address woodland issues relevant to the people. Chapter government may also experience conflicts of interest in resolv- ing problems regarding woodland issues among range units of which they are and are not a part. The Planning Committee as Community Basis The 'Nahat'a' Dziil Planning Board', which consists of two elected representatives from each range unit, and they meet with Chapter officials once a month to discuss issues and concerns from their constituent range units. Currently the plan- ning board's activities are subsidized by the ONHIR which retains an influential role in the chapter level decision making process. The plan- ners invest their time on community affairs because they receive financial compensation. Some plan- ners may not wish to assume additional responsi- bilities for developing a woodland management plan. The Planning Board members may be capable of coordinating activities with resource managers in developing woodland policies but implementing and monitoring a project is difficult for planners to under- take alone. Some planners also have indi- cated difficulties in communicating and working with local residents. Range Units as Community Basis A major concern with using range unit resi- dents as a basis for community delineation is the unequal number of homes, population, and 164 amount of woodland resource distributed among the residents of the 13 developed range units. A range unit with excess resources for community needs may wish to market surplus woodland products for income, while an adjoining range unit may not contain sufficient resources for subsistence activities. The uneven distribution of resources among range units may lead to conflict among range unit residents in the future. To mitigate this problem, management agencies need to define the amount of resources that could be harvested for ei- ther personal or commercial purposes based on biophysical condition of the each range unit. The present livestock grazing program that is organized according to range units is a positive example of cooperation to carry out woodland projects. An ex- tension forester working with range unit residents on woodland issues, could facilitate other planning activities as well. Livestock Grazing Permit Holders as Community Basis The grazing committee made up of grazing permitees is a traditional unit of social organization recognized by the Tribal and Federal rangeland management agencies. Families with grazing per- mits play an important role in maintaining Navajo tradition, family ties, and legal rights to land and resources. The livelihoods of permitees are directly affected by range management decisions. Conse- quently, grazing committee members actively par- ticipate in meetings. Since the condition of the range resource is directly affected by any wood- land management prescription, the active role of the grazing committees in resource stewardship suggests permitees as a viable community focus in taking a community based approach to woodland management. However, the community forestry literature is clear in its tenets that community based forestry programs are to improve the lives of all ru- ral people, especially the poor. Authorizing grazing committee for woodland management may add more influence to the permitees and not enough to the other residents. Non-permitee residents should receive equal opportunity to use woodland re- sources and their desires and needs should be equally considered in the planning process. Evaluation of the four levels of social organization to serve as a basis for community forestry programs suggests range units are viable community group with which to work with towards community based woodland management planning. Some of the reasons for this conclusion includes: set geographic boundaries of the range units, social and family ties facilitated by the clustered homes and neighborhood structure, existing range management program and established grazing committees for the range units, and most importantly, range units include planners, permitees, Chapter officials, — all community people. Using range units as a community basis from which to initiate a community based management planning approach will necessarily include man- agement concerns expressed by the existing or- ganizations, while allowing community residents' input regardless of their affiliation with existing organization. However, this effort to be successful will require an extension forester to live in the New Lands and work closely with the each range units' residents in establishing a woodland resource committee. Formulation of a range unit level committee will improve the residents' communica- tion abilities and will also enhance the planners' relationships with residents. This exemplifies an ideal grass-roots, bottom-up management planning approach. Information needed to evaluate the manage- ment agency's capability in restructuring institu- tional and organizational arrangements to initiate a community based approach was observed on the New Lands based on the support and cooperation provided by the ONHIR, Navajo Tribe Forestry Department, and Chapter Officials. The present management agency, the ONHIR, is willing and capable of implementing community based woodland projects. The Office of Navajo and Hopi Indian Relocation were active participants during meetings and interviews with the residents. New Lands is a part of the Navajo Nation and it is important for ONHIR to coordinate with the Na- vajo Forestry Department's woodland policies to assure compatibility. Several attempts were initi- ated as part of this research to establish coopera- tion between residents and the Navajo Forestry Department staff. Lack of familiarity with the pro- cedures required for a community based approach among traditional foresters and policy makers pos- sibly resulted in some hesitation in consulting with residents management recommendations, RECOMMENDATIONS The community based woodland management approach requires time, money, and a devoted ex- tension forester willing to learn about community residents and their interaction with woodland re- 165 sources. Classically trained forestry professionals, especially those working for rural indigenous communities, must understand and embrace the philosophy behind community based approach. The extension forester's management agency needs to mandate a policy and develop an adminis- trative structure by which the extension forester can acknowledge community resident's unique relationship with the environment and bridge the gap between management agency and the com- munity. The extension forester's task should in- clude not only providing technical support regarding biophysical concerns, but also include activities to increase community empowerment such as moral building, participation, cooperation, and basic education in woodland ecosystems. The community based approach requires foresters and land managers to assume a community supporting role, which may be a new experience for many field foresters. Working directly with local leaders and woodland users can be stressful because the field staff must give up their technocratic, authori- tarian roles in the community in favor of a role as advisor and facilitator. Agencies need to establish programs to train field foresters in dealing with lo- cal people according to ecologically and culturally appropriate modes of behavior. The effectiveness of care for land and wood- land resources by users is proportional to their sense of ownership and dependency on the re- source. For a community based management ap- proach to be successful, the people must be secure in the belief they will benefit from their investment as resource decision makers. The community based approach described in this paper is a benefit based approach where community people are helped to visualized the results of their efforts, thereby pro- viding incentives for their involvement and effort. The management agency needs to evaluate the biophysical data and explain to the community prospective benefits associated with various man- agement alternatives. Communities that use pirion- juniper woodland resources as an integral part of their traditional living can provide significant assis- tance to the sustainable management planning and monitoring activities conducted of the technical management organization. In this mutual ar- rangement between community and agency, both should agree how management task and benefits are to be shared by participants in the planning process. This study attempts to describe some of the principles and methods of community forestry applicable to communities in the United States. The community based approach to managing the New Lands pinon-juniper woodlands of the Navajo Na- tion appears desirable to the people living there and appears feasible to implement as well. The key to the success of community based woodland man- agement approach, however remains in the hands of management agencies currently with oversight responsibility for New Lands resources. The man- agement agencies working with New Lands com- munities would be well served to recognize the significance of the community based approach for the long-term sustainability of the pinon-juniper woodlands of the Southwest. REFERENCES Burch, R. William Jr. 1988. The Uses of Social Science in the Training of Professional Social Foresters. J. of World Forest Resource Management. 88 (3) No 2, 73- 109 Cernea, Michael M. 1991. Knowledge from Social Science for Development Policies and Projects. In Putting People First; Sociological Variables in Rural Develop- ment. Second Edition A World Bank Publication. 1-37. Einbender, LeGrand. 1990. Social Forestry in the Navajo Nation: Implications and Opportunities for Multire- source Management. A Thesis. School of Forestry, Northern Arizona University. Fisher, R.; Gilmore, D. 1990. Putting the Community at the Center of Community Forestry Research. Re- search Policy for Community Forestry Asia Pacific Region. Proceedings of a Seminar. Winrock Interna- tional Institute for Agricultural Development, the Ford Foundation. RECOFT Report no. 5 Bangkok, Thailand. Fisher, R.H.; Shing, K; Pandey, D.; Lang, H. 1989. The Management of Forest Resource in Rural Develop- ment: a case study of two districts of Nepal. Discus- sion Paper. International Center for Integrated Mountain Development, Kathmandu. Food and Agriculture Organization of the United Na- tions; Forestry Paper 1978. Forestry for Local Com- munity Development. Gilmore, D. A; Fisher, R. 1991. Villagers, Forests, and For- esters: The Philosophy, Process, and Practice of Community Forestry in Nepal. Sahayogy Press, Kathrnandu. Gregersen, Hans; Draper, Sydney; Elz, Dieter. 1989. People and Trees. The Role of Social Forestry in Sus- tainable Development. Economic Development Insti- tute of the World Bank. Guggenhim, Scott; Spears, John. 1991. Sociological and Environmental Dimensions of Social Forestry Project. In Putting People First; Sociological Variables in Rural Development. Second Edition A World Bank Publica- tion. 304-339. Jordan, Kenny B. Charles. 1988. Forestry Program Fights Rural Poverty: How development programs assist 166 farmers in becoming self-sufficient. J. Forestry. 88 (5): 37-41. Kelly, B. Klara. 1986. Navajo Land Use. An Ethnoar- chaeological Study. Navajo Nation Cultural Re- source Management Program. Window Rock, Arizona. Lee, G. Robert; Field, R. Donald; Burch, R. William Jr. 1990. Community and Forestry. Continuities in the Sociology of Natural Resources. Westview Press, Boulder, San Francisco, and London. Office of Navajo and Hopi Indian Relocation. November 1990. Plan Update. 167 An Overview of Woodland Projects Incorporated at Four Pueblos in New Mexico Buff Jebsen-Ross and Richard Schwab1 Abstract. — For centuries Indians throughout the Southwest have traditionally utilized and managed their Pinon-Juniper woodland resources for economic, social, cultural, and religious purposes. Basically, up to this point these uses have satisfied personal and family needs, and except for a few circumstances have not entailed intensive formal planning initiatives. However, within the past few years tribes have become increasingly aware of the development potential of their woodland resources to supplement tribal revenue, personal incomes, and employment opportunities. Accordingly, the expanded management intensity and scale has made it necessary for Indian tribes to recognize the importance of implementing comprehensive woodland management planning. There are several unique characteristics to this planning process uncommon to other land management agency practices. Differences in tribal governmental structures, social and cultural uses of the resource, land ownerships within tribal reservations, and traditional uses and perceptions to management all make planning a challenging experience. The intent of this paper is to discuss several planning initiatives coordinated by the Bureau of Indian Affairs on behalf of Indian tribes that have incorporated comprehensive project planning processes. These planning efforts have transcended into successful woodland resource development endeavors devoting special concern to sustaining the woodland resources, while supporting the traditional cultural integrity of the Indian people. INTRODUCTION The Bureau of Indian Affairs', Southern Pueblos Agency (SPA) is located in Albuquerque, New Mexico. SPA serves ten separate and distinct sovereign Indian Nations with an overall land base of 1,063,761 acres. Of these lands, 200,000 acres are considered to be commercial woodlands and an additional 150,000 woodland acres are considered to be non-commercial. The remaining reservation acreage consists of range and timberlands. All of these lands are deeded to the tribes but are placed in Federal Trust status. The scattered locations of the tribes and the large amount of forested acres complicates the management of this unique trust responsibility. Furthermore, the "multi-tribal" agency situation at SPA produces challenges and situations not found at single tribal agencies. Until recently this situation has largely ignored. Then in 1989, the Bureau of Indian Affairs (BIA) received money to initiate economic development of the woodlands resource through grant monies to the tribes. The Branch of Forestry at SPA has supported four Pueblos in applying for and receiving the grant monies from the BIA Woodlands Management Program. The purpose of these grant monies are as follows: 1. Explore the feasibility of tribal economic development through management and utilization of woodland resources, and 2. Study the potential for resource enhancement through woodland management (Schwab, 1993). This paper will discuss the planning process of these projects and grants and their application on 1 Bureau of Indian Affairs, Albuquerque, NM. 168 the ground. The planning process consists of four steps: 1. Inventorying the resource. 2. Consultation with the tribal leadership. 3. Tribal approval. 4. Project implementation. INVENTORYING THE RESOURCE The BIA provides forest inventory services to the tribes as part of its trust responsibilities. One of the elements of these inventories is determming the amount and location of commercial acreages. This in turn helps the tribes to determine the overall feasibility of economic development projects. The BIA uses a land classification scheme to determine commercial and noncommercial acreages. Land Classification Scheme For management purposes, the Albuquerque Area Office (AAO), Branch of Forestry has developed a forest classification system which divides the total forest into elements according to species composition, productive potential, accessibility, and other factors. This classification system is illustrated in Figure 1 and in the following discussion. WOODLAND Commercial woodland as defined by the BIA is land qualifying as forest, and containing less than 5% commercial timber crown cover and, diameters of at least 3.0" at diameter root crown (DRC) of woodland species or 5.0" at diameter breast height i — TIMBERLAND TOTAL FOREST— I— NON-COMMERCIAL —WOODLAND- COMMERCIAL- REGULATED (DBH) for timber species, and considered a high woodland site. Commercial woodland is a term coined to describe that portion of the woodlands producing marketable woody products which is currently or prospectively accessible, is not withdrawn from such use, and not already accounted for within commercial or non-commercial timberland (BIA 1988). The commercial forests are then divided into regulated and unregulated lands. Regulated This class includes commercial woodland which produces or has the potential to produce at least five cubic feet per acre per year of commercial volume. These areas must be accessible, not reserved, and located on slopes less that twenty percent. Non-Commercial Forest This class includes: 1. Areas administratively woodland use. withdrawn from 2. Areas that are so steep, unstable, or rocky that they cannot be used for woodland production without causing serious environmental impacts. 3. Productive areas where there is low probability during the next ten years to use the woodland because of excessive development costs. 4. Non-productive areas that lack the capacity to grow at least fifteen cubic feet of woodland per acre per year. The acreage of the Pueblo woodlands is provided in the following table (Albuquerque Area Office, Branch of Forestry, 1991): PUEBLO WOODLANDS COMMERCIAL NON-COMMERCIAL Figure 1. — Forest Type Classifications. Acoma Cochiti Isleta Jeme San Felipe Sandia Santa Ana Santo Domingo Zia SPA TOTAL 108,027 0 22,817 29,646 2,000 0 0 5,000 50.238 217,728 66,000 16,069 2,600 17,982 1 5,552 2,525 2,772 25,015 7.317 155,832 169 TRIBAL USE AND NEEDS OF THEIR WOODLAND FORESTS From these woodland areas, the tribal uses and needs has to be considered by the manager when planning development projects. Cultural Significance The woodland forests are a very significant part of the culture and ceremonies that are part of the Pueblo way of life. For example the fresh pitch of the Pinon pine (Pinus edulis) is used by "Medicine Men" for ceremonial purposes. Many of the feathers used in ceremonial dances come from birds directly associated with the Pinon-Juniper woodlands. It is a belief of the Acoma people that the spirits of deceased relatives reside in living trees within the woodland forest. This greatly complicates the management of the forest (Stanley 1992). Many of the historic sites of the early Anazazi people are found in the P/J woodlands and are held sacred. For example there are the Kow-ina ruins in Cebollita Canyon and the Sand Canyon ruins near Acoma. A Bounty of Natural Harvest The nut of the Pinon pine has been a staple of the Pueblos and many of the other Native Americans of the Southwest for many centuries. The woodlands are also an important habitat for a wide range of wildlife that has provided a historic nutritional substance for the people of the Pueblos. Hafner (1991) listed the following cpecies that are found within the woodland forest: mule and whitetail deer, antelope, elk, and rabbit. Dr. Schemnitz in his presentation in 1991 stated that quail and turkey are game birds also known to utilize the woodland habitat. Today the woodlands also provide habitat for domestic livestock A Fuel Source and Building Materials Tribal members have historically favored the juniper species of the woodland component over the pinon pine. This stems from the other more important uses of the pinon, nutrients and ceremonial. The One Seed Juniper (Juniperous monosperma) is the favored fuel source of the Jemez tribe due to its burning characteristics. With the introduction of domestic livestock and the fencing of open ranges, the need for a handy and inexpensive building material has come into play. The Rocky Mountain Juniper (Juniperous scopulorum) is favored for fence post material which is followed by the One Seed Juniper. Historically pinon was used for vigas and uprights but have been replaced with the Ponderosa pine (Pinus ponderosa). Future Demands of the Woodland Forest The Pueblos recognize the need to better manage their natural resources and thereby control their own destiny. They see a need to provide jobs and income on the reservations and see their woodland forest as a source of revenue. The tribal elders can remember open woodland forests with plenty of forage. Now these stands have closed canopies with little or no undergrowth of grasses and forbs. This is a hinderance to domestic livestock and an empty kitchen for wildlife. Through the implementation of Tribal woodland enterprises the attempt to properly manage their woodland resources is being undertaken. Sound silvicultural practices are turning overgrown stands into a productive ecosystem where man and beast can prosper. These cuts also provide employment opportunities for tribal members and a new source of revenue. DEVELOPMENT OF THE WOODLAND PROJECTS Tribal Input, the Basis For Each Project As with every other federal agency that conducts projects on the ground, the BIA must conform to NEPA standards and regulations. It is also open to outside criticism and citizen lawsuits. Unlike other federal agencies however, the BIA has a more finite and unique constituency, which is the Tribe who has sovereign rights to the land. Consequently it is the Tribe that has the final say concerning a project. Therefore, the BIA's, Branch of Forestry acts as a consultant and trustee to each of the Pueblos. 170 At SPA, the planning for each project must be fully explained to the Tribal Council. The presentation must take into account the biology of the management alternatives as well as cultural and traditional values. To facilitate this process, the Forestry Management Plan consolidates the proposed projects into one document. Some benefits of planning are as follows: 1. The Tribe takes charge of its future. 2. Gives direction to the BIA's work 3. Increases accomplishments. 4. Makes it easier to raise money. 5. Involves and educates people. 6. The plan is a basis for monitoring management activities. 7. Allows Tribe to enter in cooperative management agreements. 8. Protects precious or sensitive resources. The Southern Pueblos Agency Forest Operating Plan provides detailed descriptions about how to manage the forest resources in order to accomplish the goals of the individual Tribes. This is in accordance with the Bureau of Indian Affairs Manual (53 BIAM Supplement 2, Section 2.4) which states the following: "Forest Management Plans are required for all Indian Reservations having forest resources. The harvest of timber from any Indian forest lands will not be authorized until there have been prescribed methods of cutting based on sound suvicultural principles. The plan is a guide to ongoing forest activities and reflects the extent of the management knowledge available on the forest. Preparation of reservation forest management plans and their revisions are a responsibility of the Agency Superintendents. Tribal consultation and legally required consent will be an integral part of forest management planning. Management plans will be approved by the Area Director." The Planning Process Planning is simply the process of envisioning a desired future and developing the methods to achieve it. The process used for planning at SPA is adapted by a method developed by Dr. James Hardy (1981) for nonprofit organizations. That process uses a planning triangle seen in Figure 2. Figure 2. — Planning Triangle. This triangle consists of four components. Each component must be followed in order. Too often, organizations decided what they want to accomplish before deciding which direction they want to follow. Hardy's process helps organizations develop that direction. A description of each component is as follows: 1. The Mission Statement - This is an abstract statement of the values and purpose of the Tribe and their philosophy for the use of their natural resources. It is the focal point of the plan. 2. Long Range Goals - These are defined as a description of a desired future for the use and management of resources within a specific time frame. 3. Objectives - Every goal has one or more objectives which describes what has to happen one year from now to ensure that the goals are being achieved. 4. Action Steps - These are specific, accountable procedures which state who is responsible for accomplishing the objective and how it is going to be done. This is the only component that specifies how something is going to be accomplished rather than what is to be achieved. The successful development and implementation of a reservation resource plan involves teamwork between the Pueblo and the Bureau of Indian Affairs (BIA). In order to facilitate this, two working groups are usually formed. These are the Pueblo's Natural Resource Planning Committee and the Bureau's Interdisciplinary (ID) Team. The ID Team meets with each Pueblos' committee at the onset of the management 171 planning projects. A summary of the planning meetings are as follows: 1. Objectives of the meetings were explained to all Committee and Team members. The objectives were two fold, to learn about planning and to develop the components of the plan. 2. The benefits of planning were explained as were the planning process and planning definitions. 3. A Mission Statement was developed. This statement was usually written before the meetings by the Pueblo Governor and the facilitator and then reviewed by the committee. 4. Outside factors and the current status of all resources were reviewed. 5. Long range goals for the management of the reservation were developed by the Pueblo committee. These goals are approved by the Tribal Council. 6. Measurable and attainable objectives were developed for each goal. 7. Once the Pueblo had determined what kind of future they wanted for the use and management of their resources, action steps were written for the achievement of each goals and its objectives. 8. The Implementation Plan was written and presented at the meetings. The plan included the Mission Statement, goals and objectives, action steps, and resource summaries. The resource summaries included maps and resource descriptions. 9. The Plan was made available for review and comment and then approved by the Tribal council. 10. The plan was implemented. 11. One year from the implementation of the plan, it will again be reviewed by the ID team and the Council. The purpose of the review is to evaluate the plan's implementation, update the one year action steps, and make any necessary revisions. Organization of the Plan The Forest Operating Plan consists of three documents, as follows: 1. Forest Inventory Analysis - This report (one for each pueblo) is designed to document forest inventory field procedures, deter- mine the current forest condition, and to calculate an Allowable Annual Cut (AAC) for the forest. This report is a basic step in the forest management planning process. 2. Policies and Procedures - This document contains the general course of action that will be followed by the Branch of Forestry in order to run its program. It includes guidelines on forest management activities such as reforestation, fire management, timber sales, etc. It is divided into five sections: a. Introduction b. Goals and Objectives c. Timber Management Program d. Forest Protection e. Branch Organization Because forest policies and procedures are generally the same for each pueblo, one "generic" document was prepared which applies to all four pueblos. 3. Executive Summary and Action Plan - This document consists of a brief summary of the status of the resource and the specific projects and Action Steps which are planned for each Pueblo. The Action Steps are updated annually at the end of each fiscal year and the implementation of each Action Step implementation is tied to indi- vidual performance elements. In planning for each Pueblo, the manager must remember that each Pueblo is unique unto itself and that no two Pueblos will see things in the same light. Therefore nothing can be taken for granted. Once the tribe has settled on what it desires to accomplish, it can seek available funding to achieve its goals. For woodlands projects, they can apply for funds from the BIA's Woodlands Management program. Funding For Projects The BIA at the Agency level does not have annual budgeting set aside for woodland manage- ment. Instead monies are allocated from the Area office to fund projects as proposed by the 172 individual tribes. Four Pueblos within the jurisdic- tion of SPA have been awarded six grants from this source. Those tribes and their projects are as follows: tyyu TCT TTT A DT TCT3T (^\ r,nX/\/CDr,I A T IbLfclA rUbBLU L-LJJVlMcKL.iAL FUELWOOD PROTECT 1QQ1 1771 wnnni amd pkttft?pt?tqf 1992 ACOMA PUEBLO SPECIALTY PRODUCT GRANT 1993 SANTA ANA PUEBLO LIVE NURSERY PROJECT 1994 SANTA ANA PUEBLO NURSERY RETAIL CENTER 1994 JEMEZ PUEBLO COMMERCIAL FUELWOOD AND MULTI-PRODUCT ENTERPRISE These grants are intended to begin an ongoing enterprise that will hopefully reach self sufficiency. This is merely pilot project front money to carry the enterprise through the first year and to pay for the accumulation of inventory that is marketable revenue. This program income can then sustain the enterprise. The entity that controls the funding can either be the Tribal Administration with the support of the Tribal Council or a separate entity defined by a charter such as is found in the Pacific North West or, for example, the Mescalero Apache Timber Company. The BIA through the Agency Branch of Forestry has control of approving projects based on Tribal proposals and then holding the Tribes to the agreements of that contract. The BIA is available to advise the best means of utilizing this money to achieve the best results. Accountability of deliver- ables as stated in the grant contract is the respon- sibility of the tribal enterprise. Project Layout and Harvest It is the responsibility of the BIA to make rec- ommendations of areas to be treated. The Tribe being the land owner must designate which area they want treated and then get the approval of the Tribal Council through a resolution. Once a Tribal resolution is in hand, the Agency Forester begins work with the Tribal Administration to ascertain that preliminary work and layout meets the needs of the Tribe. An interdisciplinary team from the BIA and Tribe then meets at the proposed site to discuss concerns and any issues need to be mitigated be- fore harvesting occurs. The SPA Branch of Forestry prefers a leave tree mark on all but their Forest Pest Management treatments. With a leave tree mark you are managing the residual and the rest of the trees are excess to the crop tree needs of the stand. Thus promoting better silviculture. Once all the comments have been noted and mitigated the treatment area is turned over to the enterprise for implementation. The method of harvesting is the Tribes decision. This can be accomplished through different schemes and hiring tactics. 1. Pay individual tribal members to harvest, split, and haul materials on a load basis. 2. Hire a tribal crew to do all the work. 3. Pay individual to harvest and haul and a small crew to split sell. 4. Utilize individuals being funded through other programs to do all or part of the work 5. Subcontract the operation to tribal concerns or individuals. 6. Subcontract to outside businesses for stumpage. 7. Use a combination of any of the above. As the old saying goes, "there's more than one way to skin a cat". Problems Encountered The number one problem encountered in all but two of the Pueblos is continuity in the program. Yearly changes in the many tribal governments bring shifts in priorities. In many instances a man- ager is hired for only part of the year with no one to carry the "ball" into the next season or year. Some of the Pueblos have tried to run the entire program from a branch already overloaded with work and problems. This a sure way to failure as experienced at Isleta Pueblo. Another major problem is the trespass of the very resources designated for the project. This trespass has been conducted by tribal members and from non-members from outside the Pueblo. Once some activity is observed in the forest you can assured that someone is going to help themselves to the "easy pickings". In many cases available funding has run out before an area can be completely treated. Many 173 acres that have been prepared for harvesting are still waiting for some sort of treatment. In one instance money obtained through the sale of fuelwood products was misplaced, closing down the entire enterprise. Accomplishments and Success Stories The development of the woodland grants at SPA has enabled its staff to successfully prepare over 270 acres of commercial woodlands. An additional 210 acres have been prepared for Pest management treatment to suppress and deter the spread of two species of Dwarf mistletoe: Arceuthbium divaricatum in the Pinyon (Pinus edulis) For fiscal year 1993, the Branch of Forestry helped to obtain 30% of woodland grant monies from the Albuquerque Area Office. This was over 8% of the entire national Woodland Management budget (BIA 1994). Accomplishments by Pueblo ACOMA 1. Applied for and awarded a $ 30,000 Woodland Management grant. 2. Opened contracts with tribal members to harvest and haul 145 loads of pinon-juniper. 3. Constructed a 69,000 square foot woodyard. 4. Purchased a Supersplit log splitter from GFX Manufacturing corporation. 5. Purchased a work shed to shelter the enterprise's equipment. 6. Hired three additional tribal members to split and stack the firewood. 7. Produced 100 commercial fuelwood cords. 8. Ninety five cords sold commercially with the rest going to tribal members in need. 9. Produced $6,509 in clear profit. 10. Distributed over $26,000 in payroll to tribal members. 11. Split and stacked over 60 cords from this year's Timber Stand Improvement projects. 12. Applied for and awarded a $20,000 Specialty Woodland Product grant. 13. Purchased a Vermeer Brush chipper, model 1220. 14. Produced 450 cubic yards of chipped material. ISLETA 1. Applied for and awarded a $54,000 Woodland Grant. 2. Contracted 12 Tribal members to haul over 400 truck loads of green pinon-juniper. 3. Purchased a Duerr 20 ton and 25 ton woodsplitter. 4. Purchased a four wheeled drive pick up truck and radio equipment. 5. Provided additional salary for staff at the Tribal Cinder and Gravel Enterprise yard for scaling and book keeping services. 6. Produced over 300 cords of commercial fuelwood and sold them at $100/cord. SANTA ANA 1. Applied for and awarded a $26,400 Woodland Grant. 2. Provided $10,870 in salaries to extend the employment of the existing work force of 3 laborers. 3. Purchased one chainsaw and a hydraulic truck hoist. 4. Transplanted 294 wildlings to nursery. 5. Containerized trees were sold for a $2,000 profit. The best example of woodland management is found at Santa Ana Pueblo where they have expanded their Tribal Nursery and Greenhouse enterprise to include the harvest of wildlings in their Bosque woodlands or natural tree nursery. This harvesting is being conducted to help open dense stands of Russian olive (eleagnus angustifolia) and to reduce competition. One hundred and thirty trees have been retrieved and containerized for public sale. They plan to double that number this coming fall. Not only are they succeeding in managing their own lands, they are doing it at a profit. That profit goes to keep their employees working year round instead of seasonal. JEMEZ The project has just begun this spring and has been held up due to extreme fire danger. Even though hindered by fire restrictions the enterprise has been able to haul over 30 cords of firewood. 174 The Enterprise is proposing to cut and haul over 400 cords this year. At this time the Pueblo of Jemez is developing their Forest Enterprise that will employ 6 to 10 people year round while producing a variety of products from their forests. These will include vigas, latillas, fuelwood, fence posts, and Christmas trees. They are also planning to contract thinning and other services from other land owners. RECOMMENDATIONS AND OBSERVATIONS Another commodity of the woodland grants is the knowledge and experience gained from administering them. Some of the lessons learned are as follows: 1. The history of the area is better understood. We learn past uses as well as the environmental history. 2. The Tribe's needs and expectations are better solicited for the area. It is their land and the projects are more successful when it is something they want done. 3. The tribal constituency is educated on the biology of the prescribed management and its alternatives, either silviculturally or otherwise. Knowledge is a very powerful thing when used right. 4. Its preferred to have the Tribes plan in hand. Don't let available funding run the program. 5. The most important element is to have a positive, enthusiastic Tribal member running the operation. This person must not have other duties within the Tribal Administration. The person in this position must be dedicated to the Tribe and the resources. They must also have a great deal of initiative, be creative, and have the ability to anticipate the unexpected. BASIC SUPPORT NEEDS FOR A SUCCESSFUL OPERATION 1. WOODYARD A. FENCED AND SECURED. B. CENTRALLY LOCATED. C. EASILY ACCESSIBLE - LOADING & UNLOADING. D. LARGE ENOUGH TO HOLD STOCK WITH EXTRA ROOM TO MANEUVER EQUIPMENT AND TRUCKS. 2. WOODYARD FOREMAN 3. COMMERCIAL SPLITTER LARGE ENOUGH TO HANDLE 20" DIAMETER LOGS. 4. DELIVERY VEHICLE A. 1 TON DUALLY. B. TRAILER. OPERATIONS 1. HARVESTING A. FELLING B. LIMB AND SCATTER C. BUCKING INTO LENGTH D. LOADING In summary, given a large enough resource and a cooperative Tribal Council, there is no limits on how far an enterprise can go within the southwest. The only limits are those produced or recognized by the people within it. BIBLIOGRAPHY Hafner, David J. 1991. Non-Game Mammals of SW Pinyon-Juniper and Grassland Ecosystems. Presented at the Integrated Management of PJ and Grassland Habitats In the SW symposium. Hardy, James, 1981. Managing for Impact in Nonprofit Organizations. Schemnitz, S. D., 1991. Patterns Used by Game Birds. Presented at Integrated Management of PJ and Grassland Habitats In the SW symposium. Schwab, Beverly A., 1993. Bureau of Indian Affairs Pilot Woodlands Management Program. Managing Pinyon-Juniper Ecosystems for Sustainability and Social Needs Symposium: 146-148. USDI, BIA, 1988. Native American Woodland Resources, A National Overview. USDI BIA, 1994. FY 1993 Woodland Management Accomplishment Report. 175 The Effects of Fire on Cultural Resources Mesa Verde National Park, Colorado Kathleen Fiero1 Abstract. — In 1989 the Long Mesa fire burned approximately 3000 acres in Mesa Verde National Park, Colorado. The fire spread along the top of Long Mesa and the adjacent canyons covering an area containing approximately 200 known archeological sites. Because of the concern for cultural re- sources, all fire lines and similar earth moving tasks were done by hand with an archeologist on each crew. No heavy equipment was used. The archeological sites in the burn area varied from prehistoric to historic, iso- lated hearths to surface pueblos and cliff dwellings. A resurvey of the burn area was undertaken to evaluate the extent of fire damage. The direct ef- fects of the fire varied from very minimal impact on buried sites to complete loss for historic wood structures. Two rooms in cliff dwellings burned. During this resurvey, 23 previously unidentified sites were recorded. A project to stabilize cliff dwellings in the burn area was undertaken as was a study of the effects of fire retardant on pottery sherds. Three natural re- source studies were also funded: a vegetation map, a fire history study, and a vegetation and fuels inventory. INTRODUCTION Mesa Verde National Park is located in south- western Colorado on the northeast edge of the Colorado Plateau. The park was established in 1906, ten years before the National Park Service, to preserve and protect the cultural resources located within its boundaries. The park contains 52,000 acres and the vast majority of this area has been surveyed by an archeologist. Approximately 4,000 sites have been recorded. These sites vary from large cliff dwellings to one and two room storage units, from masonry pueblos to dirt walled pithouses. There are also wood sweat lodges, fence posts and cabins. The great majority of the sites are a manifestation of the Northern San Juan Basin Anasazi tradition and date from AD 600-1300 but there are a few earlier Archaic sites and a few that date to the historic period, AD 1860s to the present. Mesa Verde was well named by some unknown Spaniard. The flat cuesta that slopes gradually to the south is indeed green. There are three main vegetation communities in the park: pihon-juniper woodland which is dominant at the south end of 1 Archeologist, National Park Service, Mesa Verde National Park, Mesa Verde, CO. the mesa, mountain shrub at the north end of the mesa, and Douglas fir on the north facing escarp- ment. Fires have been suppressed in the park since its establishment in 1906. Consequently high fuel loads have developed in many areas of the park There have been several major fires in the park: 1934, 1959, 1972. The last major fire, the Long Mesa fire, was in 1989. THE LONG MESA FIRE The Long Mesa fire burned approximately 3000 acres. The fire started in a dense stand of pinon- juniper (fig. 1) and spread north moving into the mountain shrub community and finally burned it- self out descending the north escarpment in a Douglas fir community. The fire started on a hot, dry, windy day in early July and burned for two weeks. The effects of this fire on the park's cultural resources will be discussed in terms of the direct effects of the fire, and the effects of fire suppression activities. 176 Fire Suppression Two key management decisions had a big im- pact on the effects of the fire on cultural resources. One decision was to fight the fire by hand — no heavy equipment was used. This decision was based on previous fire experience in Mesa Verde and in other park service areas with high densities of cultural resources, a common attribute of pirion- juniper woodlands in the Southwest. The second important decision was to assign an archeologist to each fire crew Each archeologist was given a copy of the cultural resource base map for the area where they were working. Initially the archeolo- gists were sent out with each crew to flag fire lines but as the fire was being fought, this changed to flagging areas to be avoided by line work, i.e. flag- ging the sites. Also all fire fighters were told of the significance of the cultural resources and the im- portance of avoiding damage to these resources. Because all work was done by hand and because there was an archeologist on each crew, there was virtually no damage to the cultural resources by fire suppression activities. Only two sites were damaged and in both cases the damage was very minor, one upright slab was chipped and one stone was displaced from a wall. There was no vandal- ism and no artifact piles. The only graffiti noted was minor and was not in the vicinity of a site. It is worth noting that the archeologists found that keeping ahead of the fire crews, flagging sites, was easier during the fire than during mop-up when people were spread much thinner over the land- scape. Figure 1. — This photograph shows the typical condition of the pi- hon-juniper woodland after the Long Mesa fire. Note the burned pinon pine and serviceberry and the heavily spalled sandstone boulder. Figure 2. — The condition of a mesa top pueblo after the fire. The pueblo is located in the mountain shrub community. Post-Fire Survey Immediately after the fire was officially out, an archeological survey of the burn area was initiated. By overlaying a map of the burn area on the cul- tural resource base map, it was known that 194 re- corded archeological sites were in the burn area. These sites varied from hearths to large pueblos, from surface sites to cliff dwellings, from prehis- toric to historic structures. Of the 194 sites, 165 were relocated. Two of the 29 sites that were not relocated were probably consumed by the fire (wood structures). The others may have been missed because after the fire it was not easy to lo- cate metal site stakes amongst all the ash and burned trees and bushes. Also some of these sites may have been misplotted originally or damaged by road maintenance between the time of the original survey and this resurvey. Twenty-three previously unrecorded sites were located and re- corded during this survey. Most of these sites were located in areas that had been covered by a dense growth of shrub (better known as Quercus goddam- nus to archeologists) and after the fire the sites in these areas were much easier to see. Of the 188 sites evaluated for fire damage in the burn area (Eininger 1990:32), 49 (26%) were missed by the fire, 71 (38%) suffered low damage, 32 (17%) moderate damage, and 36 (19%) high damage. The 29 sites that were not relocated were not evaluated and so are not included in these totals. If any of these were totally consumed by the fire, as two may have been since they were wood structures, this would increase the percentage of sites in the high category. Of these 36 located sites suffering high damage, four were totally destroyed, and one 177 room in two different cliff dwellings burned. The four totally destroyed sites were wood sweat lodges. Pack rat debris in the two cliff dwellings caught fire and resulted in incredibly hot fires burning in those rooms. Factors involved in defin- ing low, moderate and high damage included fire intensity, vegetation loss, impact to artifacts, dam- age to rubble and amount of site area affected by the fire (ibid: 11). The high category (fig. 2) covered such observations as vegetation on the site totally burned, soil and rock oxidized, spalling of exposed stone, scorched artifacts (ibid:32). Low was used to refer to sites that were partially or wholly burned over but with little visible impact because of low fire intensity. Moderate was somewhere between these extremes. On the basis of this survey, sites were recom- mended for rehabilitation work: erosion control, water diversion, stabilization. Eighteen cliff dwellings in the burn area were stabilized in 1989 and 1990 and water diversion measures were taken as needed at these sites (Fiero 1991). Two talus slope sites were seeded. This was the only inter- vention in the burn area. No other seeding, planting or other erosion control measures were undertaken. Their was above average moisture in the fall, winter and spring of 1989/90 and vegeta- tion returned to the burn area very rapidly. Studies Undertaken as a Result of the Fire Oppelt and Oliverius (1993) studied the effects of fire retardant on sherds and concluded that it was minimal. The amount of duff around the sherd was the critical variable in determining how smoke blackened the particular sherd became. Floyd-Hanna, Romme, and Hanna (1994) com- pleted a vegetation and fuels inventory of the park, Romme and Hanna developed a vegetation map, and Floyd-Hanna and Romme (1993) studied the park's fire history by dating oak There has been no research into the effect of slurry on stone, wood, ceramics. Duncan (1990) compiled an annotated bibliography on the effect of fire on cultural re- sources. Actions Taken by the Park The park now has a fire management officer who is responsible for fire suppression activities in the park. The park policy is still to put out all fires as soon as is practical and every summer there is a helicopter and helitac crew in the park for this purpose. There is also a fuel reduction crew in the park in the summer. They reduce fuel loads around park buildings by cutting trees and shrubs and removing the wood by hand. Park buildings have been prioritized as to their importance (Research Center which houses the archeological collections first, Museum second, etc.) and fuels are reduced around buildings based on this schedule. CONCLUSIONS Those archeological sites which are the most vulnerable to damage from a wildfire (see also Romme, Floyd-Hanna, Conner 1993) are wood structures such as hogans, cabins, fence lines. Also very vulnerable to fire damage are rock art panels because of rock spalling, and cliff dwellings which contain combustible material. Less vulnerable to damage in a fire are lithic scatters and water con- trol devices such as check dams. Those sites with the least vulnerability to fire are buried unexca- vated pueblos and pithouses. As important as site type in surviving a fire, is the type of fire suppres- sion activities which take place during a fire. On the basis of the Long Mesa fire, there is no question that the lack of bulldozers and other heavy equip- ment and the use of archeologists on the fire lines reduced the amount of damage to the cultural re- sources. LITERATURE CITED Duncan, Faith L. 1990. Long Mesa Fire 1989. Fire Effects and Cultural Resources: An Annotated Bibliography. Ms on file Research Center, Mesa Verde National Park, Colorado. Eininger, Susan. 1990. Long Mesa Fire 1989: Archeologi- cal Survey and Post-Fire Assessment. Ms on file Re- search Center, Mesa Verde National Park, Colorado. Fiero, Kathleen. 1991. The 1989 Long Mesa Fire. Archeo- logical Rehabilitation, Report on Stage II: Ruin Stabi- lization. Ms on file Research Center, Mesa Verde National Park, Colorado. Floyd-Hanna, Lisa, William H. Romme. 1993. Fire Effects and Fire History of Mesa Verde National Park. Final Report. Ms on file Research Center, Mesa Verde Na- tional Park, Colorado. Floyd-Hanna, Lisa, William H. Romme, David D. Hanna. 1994. Vegetation and Fuels Inventory of Mesa Verde National Park. Ms of file Research Center, Mesa Verde National Park, Colorado. Oppelt, Norman I and Timothy Oliverius. 1993. The Ef- fects of Fire Retardant Foam on Prehistoric Potsherds. Southwest Lore, Vol. 59, No. 1, pp 26-30. Romme, William H, Lisa Floyd-Hanna and Melissa Connor. 1993. Effects of Fire on Cultural Resources at Mesa Verde NP Park Science, summer, pp 28-30. 178 Western Juniper: An Evolving Case Study in Commercialization, Ecosystem Management, and Community Development Larry Swan1 Abstract. — The talk upon which this article is based was intended to stimulate exploration and marketing of juniper and pinon products. Ulti- mately, the objectives are to better utilize fiber harvested for ecosystem management purposes, and improve local community and tribal econo- mies. The evolving case history of western juniper (Juniperus occidentalis) commercialization efforts, from an unofficial coordinator's perspective, is offered as an example of what could be done in the Southwestern United States. INTRODUCTION There are many parallels between the cultures, economies, and geography where western juniper exists (primarily Eastern Oregon, Northeastern California, and Southwestern Idaho) and the Southwestern United States. And similar to west- ern juniper, there appear to be opportunities to increase the value of fiber removed for other man- agement activities. For example, it appears that markets for pi- non-juniper novelties and high-end rustic furni- ture are wide-open. It is remarkable that none of the stores at Sky Harbor Airport, Phoenix, and few in Sedona, stock items made from Southwestern pifion-juniper species. Based on informal discus- sions with store owners, the absence of pi- non-juniper items is more a problem of supply, than demand or price. In addition, there is infra- structure already in place for the pifion-juniper fuelwood industry, and only minor adjustments would be necessary to sort the type of high-quality logs desired by wood products manufacturers from lower quality material. It is not unreasonable to assume that proceeds from the sale of pifion-juniper could be doubled, based on the market niches identified above. Real- istically, however, more work is needed before this can happen: Basic marketing research should be undertaken to confirm market niches and buyers, 1 Special Projects Coordinator, USD A Forest Service, Pacific Northwest Region, Winema National Forest (Klamath Falls, OR). and products should be tested in conditions repli- cating customer home or office environments. Following is a brief description of the process and efforts made during the last four years to commercialize western juniper. The experiences discussed should be instructive to those contem- plating ways to improve utilization and the value of fiber removed for other ecosystem management activities. WESTERN JUNIPER COMMERCIALIZATION PROJECT BACKGROUND Federal timber sales have declined drastically over the last five years in much of the Western United States. The decline in Federal timber avail- ability, as well as changing markets and improved manufacturing techniques, have contributed to economic hardship and social dislocation, espe- cially in small, timber-dependent communities. The Winema National Forest, located in South- central Oregon, about four years ago organized a "focus group" of small, medium, and large wood product manufacturers, to identify critical issues, potential areas of cooperation, and who would consider working together with the Forest Service, and other government and non-profit economic development organizations. Further impetus to discussions was provided by the shutdown of 179 several local mills: Over the course of just 18 months, 1,200 manufacturing jobs were lost out of a total regional manufacturing employment base of less than 4,000. The "focus group" met informally three times and, as might be suspected, identified "supply" as the major issue facing the industry. Follow-up interviews with manufacturers were conducted by Forest Service personnel from the Forest Products Lab (Madison), Conservation and Recycling group, and Winema National Forest, using relationships established through the "focus group". Interviews were designed to define problem areas and refine what might be possible given resources available and manufacturer interests. An updated inventory of wood products manufacturers in the region was completed at the same time (Kent 1992). Interviews yielded several partners, as well as some "doable" research projects or field manufac- turing trials. High interest was also expressed about better utilization of and markets for western juniper (Juniperus occidentalis). WESTERN JUNIPER WOODLANDS AND EXISTING INDUSTRY Western juniper woodlands occupy approxi- mately four million acres, primarily in Eastern Oregon, Northeastern California, and a small por- tion of Southwestern Idaho. It is the least-utilized wood fiber resource in this region, with an esti- mated volume of over 700 million cubic feet. About 40% of the volume is on private lands and 60% on lands managed by the Federal government (Bolsinger 1989; Gedney 1994, personal communi- cation). Historically, a small volume of western juniper has been harvested for fence posts and firewood. There are also small existing markets for juniper rustic furniture and novelties, and at least a thou- sand acres of juniper woodlands in Northeastern California have been harvested over the last couple of years for power generation biomass (Ward 1994, personal communication). WESTERN JUNIPER REMOVAL Landowners and resource managers east of the Cascade Crest are accustomed to hearing western juniper characterized as a "water sucking weed". Many have heard accounts of how old springs flowed once all the juniper was cut down. Dissent- ing voices about western juniper removal and eradication throughout its present-day range are few. Juniper removal is usually undertaken by pri- vate landowners on their own or on a cost-share basis through governmental programs, such as the "Agricultural Conservation Program" (ACP), admin- istered by the Agricultural Soil and Conservation Service (ASCS), and "Forest Service Stewardship Incentive Program" (SIP), administered by state forestry agencies. Due to lack of demand and mar- kets, juniper removed by landowners and public land managers is often piled and burnt, or simply left to decompose after being knocked down or cut. WESTERN JUNIPER COMMERCIALIZATION PROJECT A fundamental premise of the "Western Juniper Commercialization Project" is that interested manufacturers will identify and explore their own markets, if well-prepared raw material (e.g. dried and surfaced) and assistance with technical prob- lems are available. The first effort to put raw material into the hands of manufacturers consisted of slicing and drying veneer. It was thought that sliced veneer would maximize value and yield from an inherently difficult tree to mill with traditional methods and in traditional lengths. Partners were recruited and samples distrib- uted at several industry meetings. Interest stimulated by the sliced veneer sam- ples, and a local mill's willingness to experiment, led to a production run of juniper fencing material. A portion of the production run was dried in a dehumidification kiln, surfaced, and distributed to 14 different Oregon wood products manufacturers for prototype and market development. Another mill run was soon put together to test market land- scape timbers and decking. A portion of this run was dried in a steam kiln and distributed to more interested manufacturers. Early market exploration confirmed potential markets for: 1) Fencing (if material is partially dried before shipping); 2) Decking (especially higher-end radius-edge decking); and 3) Land- scape timbers. There was also strong interest in higher-graded lumber for flooring, cabinets, furni- ture, interior paneling, novelties, as well as straight substitution for species which were becoming difficult and expensive to acquire. Additionally, markets were explored and confirmed for milling residual (composites), especially if the residual was in the form of "clean" chips (less than 6% bark content). 180 STEERING COMMITTEE AND PARTNERSHIPS A "Juniper Forum" was held in 1993 to publi- cize western juniper commercial potential, biologi- cal concerns, and project accomplishments. An- other goal of the "forum" was to increase network- ing and identify more potential partners. Over 150 people attended and there are now more than 300 listings on a western juniper mailing list main- tained by an industry partner. Membership on the "Steering Committee" is about equally divided between industry, govern- ment/non-profit, and landowners. Membership mix was designed to reflect the wide range of interests potentially involved in commercializing a new species. The "Steering Committee" gets to- gether about four times per year at locations around the state, and meetings are open to all interested parties. Committee members recently toured the eastern red cedar industry in Missouri to learn about and compare techniques and mar- kets for a species closely-related to juniper: Eastern red cedar (Juniperus virginiana). The committee has helped the unofficial Forest Service coordinator define and prioritize proposed projects, and find partners and funding to make them happen. Projects are based on problems identified by industry, scientists, and other coop- erators, and range from improving inventory in- formation (to include more explicit ecosys- tem-related data categories) to solid wood drying trials. Partnerships are necessary to make things hap- pen, and they change according to project re- quirements and organizational or business interest. A common thread though, is that there is reliance on government, industry, and non-profit organiza- tions and sources, rather than just one or another. Over 40 different businesses, institutions, or- ganizations, and agencies have actively partici- pated in western juniper projects. This includes at least 25 different private firms, five research insti- tutions (Oregon State University, University of Montana, Forest Products Lab, and Pacific NW and Intermountain Forest and Range Experiment Sta- tions), three state forestry agencies (Oregon, Cali- fornia, and Missouri), Oregon Economic Develop- ment Department, three local county economic development and two Resource Development and Conservation (RC&D) organizations, as well as 10 Bureau of Land Management, Extension, and National Forest System offices in Oregon and California. Critical logistic and financial support has been provided through the Community and Rural Development programs, Forest Service Pa- cific Northwest Region, State and Private branch. JUNIPER WOODLAND ECOLOGY AND RESEARCH The ecological side of the "commercialization process" is not being ignored. Over a year ago, a position paper was put together for the "Eastside EIS" Team, mandated by President Clinton in 1993 as part of the President's "Forest Plan" (Pacific Northwest Region) (Swan 1994). Prominent field scientists from different institutions helped with this effort and, in part due to Steering Committee efforts and field scientist input, western juniper woodlands will be specifically addressed in this major policy document. Contacts between the Forest Service coordina- tor and field scientists active in western juniper woodland research continue: A field demonstra- tion area is being developed on private land to showcase the latest in harvest techniques and juniper management research, training sessions for key personnel who will have to administer increased juniper harvests on private lands are planned, and a work plan and budget are being investigated to better inventory key juniper woodland ecological attributes and commercial potential. PRODUCTION AND FIELD TRIALS, AND MARKETING PROJECTS Production and field trials involving juniper have lead to preliminary reports and descriptions for a variety of topics, as well as better definition of potential products and markets: 1. Slicing and drying; 2. Scaling methods comparisons; 3. Composites; 4. Interior stains and finishes; 5. Bending properties; 6. Pallets; 7. Fencing lumber recovery; 8. Harvest machinery comparisons; 9. Pellets and BTU; 10. Fasteners; 11. Bark as cement aggregate; 181 12. ShrirLk/swell properties of solid-wood pan- els. Production and field trials in progress include: 1. Drying schedules; 2. Edge-glued panel working environment comparisons; 3. Moisture meter correction factors; 4. Furniture and cabinet parts; 5. Flooring; 6. Exterior stains and finishes; 7. Peeling characteristics and veneer markets. Planned production and field trials, and mar- keting projects include: 1. Debarking; 2. Harvest equipment and techniques; 3. Log storage; 4. Production-level kiln schedules; 5. Another juniper forum; 6. Rustic furniture market niche ID and de- sign; 7. Shavings; 8. Buyers/sellers data base; 9. Harvest and management demo area; 10. Newsletter; 11. Mulch; 12. Further bending product/market explora- tion. OBSERVATIONS OF AN UNOFFICIAL COORDINATOR Markets for western juniper are confirmed and industry interest is increasing. Interest will no doubt grow even faster as technical and economic problems are addressed. Following are observations from the perspective of an unofficial coordinator about what has worked and not worked, which may be valuable to others contemplating commercializing "under- appreciated" fiber resources: 1. Stimulating Market Dynamics - Well-pre- pared (e.g. dried and surfaced), raw material has to be available to manufacturers for prod- uct and market identification and exploration. New markets and niches are constantly being identified through this process. 2. Ecology - Ecological issues, such as water- sheds and social systems, are directly af- fected by commercialization and should be addressed at the same time. Efforts can be coordinated and sometimes combined. For example, a better inventory of western ju- niper woodlands is needed for both eco- logical and commercial purposes. Based on discussions with scientists and industry, some data categories are similar and can be combined. 3. Identification of Unique Characteristics - It is important to identify a species' unique characteristics to assist marketing efforts, as well as dispel stereotypes. This may take research assistance, both in the lab and in field trials with manufacturers. 4. "Eggs in One Basket" - It is advisable to di- vide efforts between cottage-level manufac- turers and larger commercial producers. Different markets and products will be identified. 5. Problem/Project Definition - A critical aspect of private, non-profit, and government coop- erative projects is "problem and project defi- nition". Government entities and industry need to do their homework about local re- sources before deciding there really is a prob- lem and a need to involve additional outside organizations. A "team" approach improves problem and project definition. Team mem- bers should include representatives from en- tities with direct interest in a project's out- come, such as industry, local economic devel- opment organizations, research, government, and private landowners. 6. "Doable" Projects - Ensure that when proj- ects are defined with industry and other af- fected interests, at least some are "doable" in the short-term, and have clear economic benefit and a reasonable chance of success. "Nothing begets success like success." 7. Multiple Primary Producers - More than one primary producer is needed before most manufacturers will commit to prod- ucts and markets. Reliability is a bigger is- sue than size and volume. Sole sourcing is risky at this stage of the industry. 8. Project Coordinator - Someone needs to or- ganize and coordinate. In the early stages, L government can contribute by helping to identify and combine different resources in the public and private sector. 9. Communication - Some form of regular communication is necessary to help things along and recruit more partners - newslet- ters, open forums, field trips, and meetings all help. 10. Steering Committee Formation and Com- position - An informal "Steering Commit- tee" or similar body is important to identify problems, prioritize projects, network, gain funding support, and look into the future. A conscious decision was made to integrate membership of the "Western Juniper Steering Committee" to reflect various in- terests, ranging from landowners to public land managers, and manufacturers to eco- nomic development organizations. 11. Commitment and Time Line - Commer- cialization of a new fiber resource is not a one-year, one-grant process. It takes consis- tent leadership, intense communication and networking, and constant refinement of which problems will receive highest prior- ity. 12. Federal or State Agency Support - If gov- ernment agencies want to have influence in how or what occurs, they have to find ways to assist and where assistance will be ac- cepted. In the case of western juniper, the Forest Service has played a key role because of current legislation and funding pro- grams, as well as the unique multi-faceted nature of its mission (which includes Re- search, State and Private, and National For- est system). Other organizations may play a similar role in other situations. 13. Funding and Partnerships - In addition to a "coordinator", funding is necessary for op- erations. As little as a couple of hundred dollars is often the difference between making something happen or a project fal- ling apart. "In-kind" services or supplies are also necessary, but not feasible for eve- rything. Partnership projects can become so logistically complex that minor problems cause major repercussions. REFERENCES Bolsinger, Charles L. 1989. California's western juniper and pinyon-juniper woodlands: Area, stand charac- teristics, wood volumes, and fence posts. Resource Bulletin PNW-RB-166. Portland, OR: U.S. Depart- ment of Agriculture, Forest Service, Pacific Northwest Research Station. Gedney, Donald 1994. Personal communication. Port- land, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. Kent, Peter 1992. Forest products industry review of the Klamath Basin. Unpublished manuscript. Prepared for the U.S. Department of Agriculture, Forest Serv- ice, Winema National Forest, Klamath Falls, OR. Swan, Larry 1994. Western juniper management and commercialization: An emerging issue for Eastside EIS. Unpublished manuscript. Winema National Forest, Klamath Falls, OR. Ward, Barney 1994. Personal communication. Alturas, CA: California Department of Forestry. 183 Tres Piedras Pinon-juniper Silviculture: A Partnership Project Between the USDA Forest Service and New Mexico State University John T. Harrington1, Jim Fitch2, and Patrick A. Glass3 Abstract. — In 1993, a Partnership Agreement was made between the Car- son National Forest and New Mexico State University to examine the effects of thinning intensity on pinon-juniper woodland ecology on the Tres Piedras Ranger District. The goal of the project is to evaluate fuelwood harvesting effects on the regeneration of overstory species, wood production, and herbaceous ground cover. The project area had been harvested in large blocks, 60 to 90 ha, from 1979 through 1990. Preliminary surveys of the area indicate that these blocks were relatively uniform in initial basal area and edaphic conditions. This area provides an ideal laboratory for evaluat- ing the effects of thinning intensity on the ecological processes on pinon- juniper ecosystems. INTRODUCTION The pinon-juniper woodlands of the south- western United States have, and continue to satisfy many of man's needs. This utilization occurred well before written records existed, with estimates of up to 20,000 years before present (Buckman and Wolters 1987). Use of the pinon-juniper woodlands includes the harvesting of tangible products as well as satisfying religious, spiritual and cultural needs (Miller and Albert 1993). Commonly harvested products include; products from the trees them- selves such as fuelwood, fence posts, pinon nuts and juniper berries. Other products harvested from the pinon-juniper woodlands include; wild- life, forage for livestock and water via the water- sheds encompassed by the woodlands. Non- material uses of the pinon-juniper woodlands include recreation, including wildlife watching and hiking, and the spiritual and cultural associations many Native American groups have with these woodlands. The pinon-juniper woodlands of the southwest are fast becoming recognized as one of the regions more valuable natural assets. The pressure on these woodlands for the products and uses stated above is increasing with the increasing population of the southwest. As this pressure escalates, the likelihood of a long-term disruption in the administration of the woodlands also escalates. This is especially true for the pinon-juniper woodlands administered by public land agencies who need to address the needs of many, sometimes divergent, user groups. The U.S.D.A. Forest Service has been developing sustainable, multi-resource management strategies for pinon-juniper woodlands to address these changing needs (Buckman and Wolters 1987, Tidwell 1987). Administration of these woodlands can be perplexing. The relatively slow and poorly understood ecosystem dynamics of the pirion- juniper woodlands increase the difficulty of evaluating an administrative decision with regard to the ecosystem health of these woodlands. This point necessitates that the administration of pinon- juniper woodlands be based on an understanding of the biological and ecological processes of the woodlands. However, studying an ecosystem and 1 Assistant Professor of Tree Physiology and Superintendent of the Mora Research Center, New Mexico State University. 2 SiMculturist, USDA Forest Service, Carson National Forest. Senior Research Assistant, Mora Research Center, New Mexico State University. 184 generating meaningful information to assist land managers can be quite costly. Buckman and Wolters (1987) noted that direct Federal funding for such studies is somewhat restricted. Understanding such restrictions and recognizing the need for such information on the pihon-juniper woodlands, the Carson National Forest and New Mexico State University entered into a Partnership Agreement in 1993. The Partnership Agreement provided a mechanism for the two organizations to collaborate without the exchange of funds. THE PARTNERSHIP AGREEMENT Project Location The project site is located on the Tres Piedras Ranger District of the Carson National Forest in northern New Mexico. Specifically, the project area is south of the town of Tres Piedras, NM (Figure 1). The Comanche Rim is the western border of the project area and U.S. Highway 285 is the eastern boundary with the exception of the 1990 thinning unit which is immediately east of the highway (Figure 2). TRES PIEDRES, NM PROJECT AREA • TAOS, NM • MORA, NM • SANTA FE, NM ALBUQUERQUE, NM • LAS CRUCES, NM A Figure 1. — Location map of the Tres Piedras pihon-juniper project area. Site History There are approximately 4,035,340 ha of pinon- juniper woodlands in New Mexico (Fowler et al. 1984). The Carson National Forest administers approximately 135,970 ha of pihon-juniper wood- lands (CNF Plan Amend 7, 1990). Approximately 16% of the pinon-juniper woodlands administered by the Carson National Forest are within the Tres Piedras Ranger District. There is extensive evidence of man's presence in the project area. Archeological surveys of the area indicate that hunter-gathers initially used the area during the Archaic Period (5,000 - 0 B.C.) with lithic scatters being the predominate form of archeological site (Elder 1994). European man's arrival in the area appears to be associated with the arrival of the southern spur of the Denver & Rio Grande Railroad in the 1880's (Carpenter 1994). A good review of this period can be found in Logging on the Denver and Rio Grande by Chappell. There is also evidence of wood cutting activity in the past as indicated by old, large stumps from trees felled by axes. Using Budy and Meeuwig's (1987) classification, these stands would be classified as old-growth, high-graded stands. 185 In 1979, the Tres Piedras Ranger District and the Carson National Forest initiated a program of thinnings in the project area. The initial prescrip- tion was to thin the stands to a residual spacing ranging from 6 x 6 m to 12 x 12 m (20' x 20' to 40' x 40'). Leave trees were chosen primarily on the criteria of form. The intent to generate an uneven- aged or even-aged condition is ambiguous, how- ever, this approach closely resembles Bassett's (1987) classification of a 2-step shelterwood system. The thinnings which were imposed were not initially designed as a research study, but rather as a management prescription. The trees were marked for removal by several different Forest Service personnel over the time of the thinnings. The trees were felled and removed by the general public, predominately in the form of fuelwood. Many of the marked trees however, were not re- moved. This combination of different marking personnel and variable tree removal resulted in a wide variation of thinning intensities imposed on the project area. Also, the high frequency of archeo- logical sites in the project area left many areas undisturbed by the thinning activity. These undis- turbed areas, of variable sizes, could possibly sat- isfy the needs of control plots depending on their size. The project area was also under an active grazing program prior to beginning the thinnings and throughout the thinning treatments. The project area continues to be grazed by livestock today. Project History In the Spring of 1992, the Carson National For- est Silviculturist, Jim Fitch and myself (John Har- rington) visited the project area to discuss pinon- juniper silviculture alternatives for the Carson National Forest. Following several additional visits to the project area, we concluded that post- treatment measurements of the thinning treat- ments were needed in order to assess the effects of the treatments. There was enough variability in thinning intensity during each year's treatment that we felt useful information could be generated regarding thinning intensity effects on the wood- lands. In mid-summer of 1992 we met with the Tres Piedras District Ranger and staff to discuss alternatives to evaluate the thinning treatments. It was concluded during these meetings that the likelihood of independent funding for the project would be limited. However, by combining the two organizations resources we could achieve the mutual objectives of both organizations. Over the next nine months a Partnership Agreement was drafted and submitted for approval by New Mexico State University and the Carson National Forest. On October 20, 1993, with the signature of the Carson National Forest Supervisor, the Tres Piedras Pihon-Juniper Partnership Agreement was approved. Project Goals and Objectives The Partnership Agreement set four broad goals to be achieved. These goals were: 1. To develop a database on the pinon-juniper woodlands in the project area. 2. To examine the impacts of thinning inten- sity on ecological processes in the pinon- juniper woodlands. 3. To develop an efficient, and thorough plot inventory procedure that can supply infor- mation to assist in administrative land management planning; to help create a da- tabase of ecological and biological informa- tion to be used in continuing research en- deavors. 4. The establishment of permanent plots in the pinon-juniper woodlands on the Carson National Forest, Tres Piedras Ranger Dis- trict. The situation on the project area lent itself well to a replicated treatment study over time. However, to ensure there is in fact replication of treatments over time, several assumptions needed to be made and validated. These assumptions were: 1. The different thinning years were on simi- lar sites with similar initial site and stand conditions. These similarities include simi- lar edaphic and climatic properties as well as similar stand structures prior to thinning treatments; (i.e. the treatments were im- posed on one initial population). 2. Each thinning year had similar variability in thinning intensity; (i.e. replication of treat- ments over time). 3. There was sufficient magnitude in thinning intensity whereas replicate plots of each thinning intensity could be located in each thinning year; (i.e. sufficient within year replication of treatments). Project Structure The project is divided into two phases. The first phase involves the validation of the assumptions, 186 obtaining preliminary wood fiber production effects of the thinning and obtaining the informa- tion necessary to design the second phase of the project. The second phase involves the bulk of the work towards the attainment of the main goals of the project. Phase 1 Objectives The specific objectives of the first phase of the project are: 1. To determine the magnitude and range of thinning intensity (based on ground line basal area) on the project area. 2. To provide variability information necessary to design the second phase of the project area. 3. To locate plots for use in the second phase of the project. 4. To generate preliminary data on the diame- ter growth response of residual trees. Materials and Methods Thinning Unit Determination— The project area had annual thinning tracts from 1979 through 1990. Each tract was cut during the late summer and fall (mid August through September). An exception to this was part of the 1988 thinning tract which was re-entered in 1989. Thinning years 1979, 1982, 1985, 1990 and the portion of the 1988 thin- ning tract not re-entered in 1989 were selected for initial inclusion in the project. A minimum of 25 plots were to be established in each of the five selected thinning years. Thinning area boundaries were determined using aerial photograph interpretation from the 1990 air survey of the area and from records lo- cated at the Tres Piedras Ranger District office. Thinning unit boundaries were delineated on acetate overlays of the aerial photographs (scale = 1:12,000) and areas calculated using a Summa- graphics digitizing table. Following area size determination, plot loca- tions were determined by randomly dropping a dot grid over the 1990 aerial photographs of each thinning unit. The distance between each dot was 5 chains. To sufficiently cover each thinning unit and satisfy the minimum number of plots for each tninning unit, a distance of 10 chains was used within and between plot transects. 2 Plot Procedure — At each plot center a 202 m (1720th acre) fixed radius plot was established. Plot center was marked initially with a wooden stake. (During subsequent visits to the plots, plot center was reestablished with 0.7 m steel marker and a plot tree was tagged using a 5 cm aluminum tag). The fixed radius plot had a radius of 8.02 m (26.3 feet). Species and ground line diameter was recorded for all standing trees and stumps greater than 2.5 cm (1") within the plot boundary. Ground line diameters were measured using tree calipers. Also, all advanced regeneration (possessing secondary needles but less than 5 cm ground line diameter) was recorded for the plot. Estimation of the num- ber of seedlings not considered as advanced regen- eration was also recorded. Wood cores were taken from two overstory pinon trees at each plot. Cores were taken at a height of 1.37 m (4.5 feet) from the ground. Cores were placed in plastic drinking straws labeled with the appropriate plot information. Trees for coring were selected using the following system. The first pinon encountered proceeding clockwise from the north radius and the first pinon encountered proceeding clockwise from the south radius. Data Analysis — Ground line diameter data was used to determine current plot basal area and esti- mate the basal area removed in the tHnning treat- ment. These two values were combined to estimate the initial basal area prior to the thinning treatments. Graphic analysis and non-parametric analysis proce- dures (Wilcoxon Rank Sums and Kruskal-Wallis Test) were used to compare the distributions from each thinning year of the initial basal areas and thinning treatment intensities. All analysis were performed using the SAS 6.03 statistical analysis software pro- grams (SAS Institute, Cary, NQ. Tree cores were analyzed in the laboratory to de- termine tree age and annual increment from 1969 to the present (fall/winter 1993). Cores were extracted from the plastic straws and stained using a phloro- glucinol staining procedure as outlined by Patterson (1959). Annual ring widths were measured for the previous 25 growing seasons (1969 through 1993) to provide reference contrasts from site to site with regard to diameter growth rates. Ring widths were measured under a binocular dissecting microscope using a 1.2 magnification setting with 10X ocular lenses. One standard ocular lens was replaced with a 10X calibrated ocular. The calibrated lens allowed for the determination of ring widths to the nearest 0.1mm 187 PRELIMINARY RESULTS The 202 m2 acre plots of the first phase have been established. Analysis of the data from these plots is still being conducted. Initial evaluation of initial basal area indicate that the 1988 thinning unit was not similar with regards to basal area prior to the thinning treatment. Both non-parametric procedures used to 7 1 e 10 1990 70 110 130 100 170 180 210 ' 1 ' ' 1 10 . —1- — 30 t i i 80 _J_1 * 1 ' ' 70 1 1 1 a 00 1 It 1 1 110 1 1 190 1 1 100 170 1 1 190 210 1985 10 30 80 70 00 110 130 180 170 180 a to e S. m V C « ?• ^ 2 o 1982 1M l_ 10 30 90 70 90 115 130 190 170 190 210 73 E < I3 o 10 1979 — I , i t I I I- I I 1 1 t 1 1 \ 1 30 90 70 190 170 190 210 90 110 130 Basal Area CI ass Figure 3. — Calculated initial ground line basal areas of the thinning units. Values on X-axis indicate the median of the bar range. The 1988 thinning units has 13 uncut control plots. 188 evaluate the initial basal area data indicated the 1988 thinning unit did not have the same distribution as the other four thinning units. The initial basal area of the 1988 thinning unit ranged from 10 ft2/ac to 110 ft /ac (Figure 3). The initial basal area for the remain- ing four thinning years being evaluated, 1979, 1982, 1985, and 1990, ranged from 20 ft2/ac to 200 ft2/ac (Figure 3). r 1982 F 1 1 1 s. I 4 - Hltt m Paioantaga Gata Con Figure 4.— Percent of initial basal area removed by plot frequency for the 1979, 1982, 1985, 1988, and 1990 thinning units. 189 The non-parametric analysis of the percent of initial basal area removed (tliinning intensity) also indicated that the 1988 tliinning year was not representative of the other 4 thinning years being examined. This discrepancy of the 1988 harvest can also be seen graphically. Only 2 of the 25 plots measured had greater than 30% of the initial basal area removed (Figure 4). The other 4 tliinning years had a much greater frequency of heavier harvests (Figure 4). Phase 2 Objectives The objectives of the second phase of the proj- ect are less defined at this point. Specific objectives will not be developed until the analysis of the first phase data is complete. However, several areas which will be addressed have been defined. These areas include: 1. Examination and evaluation of spatial and temporal trends in regeneration conditions involving overs tory species. 2. Examination and evaluation of edaphic and microsite factors impacting regeneration of overstory species. 3. Examination and evaluation of the diversity and abundance of herbaceous ground cover. 4. Examination and evaluation of the wood fi- ber response in residual overstory species. 5. To provide a thorough inventory (database) of species present and their distribution on the sites for the establishment of long-term studies. Materials and Methods Exact plot procedures have not yet been de- termined for this phase of the project. However, it will involve a combination of fixed size plot and variable size plot techniques to satisfy the objec- tives of this phase. SUMMARY The work thus far indicates that the 1988 thin- ning year will not be part of the replication study. Further evaluation of the age and diameter growth data will provide further information on whether the four remaining thinnings units can be consid- ered as originating from a single population. Also, the analysis of the tree ring data will provide some preliminary information as to the effects of the first entry removals on residual trees. Completion of the first phase should be done by November 1994. Field work should begin on the second phase early in the Spring of 1995. ACKNOWLEDGMENTS The authors of this manuscript would like thank several individuals for their assistance in this project. These people include Dan Rael, District Forester on the Tres Piedras Ranger District; Jack Carpenter, Forester on the Tres Piedras Ranger District; Henry Sandoval, Forestry Technician on the Tres Piedras Ranger District; Brett Coleman, Range Conservationist on the Tres Piedras Ranger District; and, Andy Lindquist, former Carson Na- tional Forest Supervisor and extremely helpful volunteer on this project. Thanks. LITERATURE CITED Bassett, R.L. 1987. Silvicultural systems for pinyon- juniper. In: R.L. Everett (ed.) Proc. Pinyon-Juniper Conference, Jan. 13-16, 1986, Reno, NV. pp. 273 - 279. Buckman, R.E. and G.L. Wolters. 1987. Multi-resource management of pinyon-juniper woodlands. In: R.L. Everett (ed.) Proc. Pinyon-Juniper Conference, Jan. 13-16, 1986, Reno, NV. pp. 2 - 4. Budy, J.D., and R.O. Meeuwig 1987. Pinyon-juniper silvics and silviculture. In: R.L. Everett (ed.) Proc. Pinyon-Juniper Conference, Jan. 13-16, 1986, Reno, NV. pp. 244 - 248. Carpenter, J. 1994. Personal communication. CNF Plan Amend 7,. 1990. Carson National Forest Plan, Amendment 7. Elder, M.A. 1994. Personal Communication. Fowler, J.M., R.D. Bowe, B.E. Peacock, K.C. McDaniel, T. Garner-Hurt, J.R. Gray, and M. Cardenas. 1984. Wood fiber production on pinyon-juniper woodlands in New Mexico. NMSU Agric. Exp. Sta. Res. Rep. No. 519 32p. Miller, R.K. and S.K. Albert. 1993. Zuni cultural relation- ships to pinon-juniper woodlands. In: E.F. Aldon and D.W. Shaw (eds.) Proc. Managing Pirion-Juniper Eco- systems for Sustainability and Social Needs, April 26- 30, 1993, Santa Fe, NM. pp. 74-78. Patterson, 1959. Distinguishing annual rings in diffuse porous tree species. Jour, of For. 57: 132. Tidwell, D.P 1987. Multi-resource mangement of pinyon- juniper woodlands: times have changed, but do we know it? In: R.L. Everett (ed.) Proc. Pinyon-juniper Conference, Jan. 13-16, 1986, Reno, NV. pp. 5 - 8. 190 Pinon Pine Seed Production, Collection, and Storage Richard M. Jeffers1 Abstract. — Key phases in the seed production cycle of pinon are summa- rized. Seed production is cyclic and good crops occur at 2-7 year intervals. During the 58-year period, 1936 to 1994, good seed crops were produced on the average every 4.1 years. Seed usually becomes mature and collectible in mid-September, which is about 26 months after the start of the seed produc- tion process. Seed yields on individual trees may exceed 20 pounds. Good seed stands produce an average of about 250 pounds of seed per acre in good seed years. Seed yields per acre can be increased through careful se- lection and retention of the best and most consistent seed producers, and elimination of poor seed producers by thinning in selected high yielding stands. Seed production can be further increased by transplanting good seed producers into high yielding stands and/or establishment of seed or- chards. Seed orchards can be established by transplanting seed bearing trees selected on the basis of heavy and consistent seed production (transplant orchards), or with seedlings produced from seed collected from selected heavy seed producers (seedling seed orchards). Currently, most pinon seed is collected by manually picking seed from the ground. Collec- tion of seed from individual trees in enhanced seed stands could be mecha- nized. In seed orchards seed collection can be highly mechanized through use of a net retrieval system such as used in southeastern pine seed or- chards. Good quality pinon seed can be stored for 5-10 years or more when stored in sealed containers, at 5-10 percent moisture content, and at 0-20 degrees F. A recommendation is made for the establishment of a complete pinon seed enterprise which would include seed production, collection, treatment, storage and marketing. INTRODUCTION Probably the most important cash crop pro- duced by pinon (Pinus edulis Engelm.) is the nigh quality, edible seed (nut) crop. There is a very high demand for pirion nuts because of their excellent taste and high nutritional value. Pirion seeds are sold as edible nuts and are used in candies, cookies, and other home and restaurant foods. In addition, pirion seed is also used for production of seedlings for landscaping purposes in the Southwest. While pirion nuts have been a food staple by the American Indian in the Southwest for many centuries, and a significant portion of the nut crops remain in the Southwest, the majority of pirion nuts are shipped to large eastern cities, especially New York City (Lanner 1981). The annual demand for pine nuts of all species (shelled and unshelled) exceeds 6 million pounds. In 1991 and 1992 it was estimated that as much as 7 million pounds of pirion nuts (unshelled) were harvested and an additional 5.3 million pounds of shelled pine nuts were imported (Delco et al. 1993) for a total of 12.3 million pounds of nuts for both years for an average of 6.16 million pounds per year. Annual imports of pine nuts into the U.S. appear to be inversely related to the availability of pirion nuts. Delco et al. (1993) reported that shelled pine nut imports declined from 4.0 million pounds in 1989, a poor pirion seed year, to 2.6 million pounds in 1992, which was a good pirion seed year. During this same period the import price per pound of shelled pine nuts increased from $2.19 in 1989 to $4.69 in 1992. And, it appears ' Regional Geneticist, USDA Forest Service, Southwestern Region, Albuquerque, NM. 191 that pirion nuts are not currently being offered to the food industry at prices that are competitive with these import prices. It seems logical then, that any increase in the annual supply of pirion nuts would be of substan- tial benefit to the pirion nut industry and, in turn, would provide added income to the people of the Southwest. Increases in the annual supply of pirion nuts, at more competitive prices, can be achieved relatively easily by: 1) harvesting more seed during good seed years, 2) increasing seed yields per tree by selection of consistent high seed yielders, 3) increasing seed yields per acre by increasing the number of high seed yielders per acre, 4) using cultural techniques to increase seed yields per tree, 5) improving seed collection efficiency through mechanization, 6) use of long-term (up to 10 years) seed storage, and 7) development and utilization of a pirion nut marketing strategy. Basic information on pirion seed production, collection and storage is provided in this paper that can be used in the development of a strategy for the establishment of a pirion seed enterprise, which would include all phases of a seed business including seed production, collection, storage, packaging, and marketing. SEED PRODUCTION Seed Production Cycle For most north temperate coniferous species, the seed production cycle (the entire time period between initiation of reproductive initials and mature seed is produced) takes place over two consecutive growing seasons (2-year cycle) for species such as Douglas-fir, the spruces, and true firs, or three growing seasons (3-year cycle) for most pine species, including all of the pirion species (fig. 1). DOUGLAS -FIR, SPRUCES, TRUE FIRS FIRST YEAR SECOND YEAR THIRD YEAR J A S 0 N D J F M A M J J A S 0 N D A F M A M J J A S 0 N D Reproductive Buds Differentiate and Develop I I I I I I I I I | I | | | | | | | | | | | | | Dormant | | | Pollen Development | | | Pollination I | | Ovule Development | | j Fertilization | | | | | | | | | Cones/Seed Develop | | Seed Mature I I I I Seed Fall PINES FIRST YEAR J A S 0 N D SECOND YEAR JFMAMJJASOND THIRD YEAR AFMAMJJASOND | | | | Reproductive Buds Differentiate and Develop MINIM! II | | | | | | | | | | || 1 Dormant MM Pollen Development | | | Male and Female Strobili Emerge from Buds MM Pollination | | || || 1 1 Ovule and Conelet Development || 1 II 1 1 1 1 II II 1 Conelets Dormant || | M 1 1 Conelets Resume Growth | | Fertilization | | || || | | Cones/Seed Develop || 1 | Seed Mature | || | Seed Fall Figure 1. — Cone and seed production cycles of Southwest conifers. 192 In the Southwest, male and female reproduc- tive structures in the conifers are initiated in late summer and mature male and female reproductive structures (strobili) emerge from reproductive buds the following spring. In the 2-year cycle species, the remainder of the seed production cycle from pollination to seed maturity, occurs between May and September of the second growing season. In contrast to the 2-year cycle species, most of the pine species, including pinon, require another full year for mature seed production to occur. In South- western pines, male and female reproductive structures are initiated in late August and Septem- ber, after height growth has been completed, of the first growing season. Growth of these initials is completed by October. These structures remain essentially dormant until they emerge from the reproductive buds in late May to early June of the second growing season. This stage is followed shortly thereafter by pollination. At this stage, mature pollen grains are shed and disseminated by wind to females, usually on adjacent trees. After reaching the females the pollen grains germinate and form pollen tubes. The pollen tubes and ovules, which later develop into seed, then start to rapidly develop; but their growth is soon arrested and these structures remain essentially in a resting condition throughout the remainder of the second growing season and second winter. Growth of the dormant females resumes about one year after pollination in late May to June of the third growing season. Fertilization, the fusion of a single sperm cell (male) and an egg cell (female), then occurs in late June to early July of the third growing season. The fertilized egg then divides and differentiates to form the embryo. Soon after fertilization the cone and enclosed embryos rapidly develop and the cones and seed become mature in late August to September of the third growing season. These key phases in the seed production cycle of pinon, first described by Little (1938a and b), are summarized in Table 1. Cone and Seed Crop Frequency Cone and seed crops in the conifers do not occur on a regular annual basis; they occur periodically. This annual variation in cone crop production is related to and affected by the biological characteristics of individual species, internal nutrient supplies, and by external conditions such as weather, insects, diseases, and predation by birds and mammals (Eremko et al. 1989). Intervals between good cone crops may be Table 1. — Key phases in the seed production cycle of pinon 1. Pinus edulis Engelm First Year 1 . August 1 5 - September 30 Buds containing male and female primordia differentiate and develop. Second Year 2. May 15 -June 15 Male and female reproductive structures (strobili) emerge from buds. 3. June 15 - June 30 Pollination occurs. 4. June 15 - August 31 Rapid conelet and ovule (develop into seed) development. 5. August 15- December 31 Conelets dormant. Third Year 6. May 1 - May 1 5 Conelets resume growth. 7. June 21 - July 7 Fertilization occurs. 8. July 1 - August 31 Rapid cone and seed development. 9. September 1 - September 15 Seed mature. 1 0. September 7 - October 31 Seed fall. 7 After Little 1938a,b and Ronco 1990 as little as 2 years as in jack pine, lodgepole pine, and Scots pine; 3-4 years in the spruces and Douglas-fir; or every 4-5 years for other pines, such as the white pines, ponderosa pine, and pinon in the Southwest. Good seed crops in all southwestern pine species tend to be synchronous, occurring in the same year over large geographic areas; that is, when a good seed crop occurs in one pine species then there are usually good seed crops in the other pine species throughout the Southwest. In pinon, good seed crops usually occur every 4 to 7 years over the entire pinon area in New Mexico, eastern Arizona and southern Colorado (Barger and Ffolliott 1972). During the 58-year period, 1936 to 1994, reports in the literature (Little 1941, Barger and Ffolliott 1972, Betancourt et al. 1993) and seed collection records showed that good pinon nut crops occurred throughout the Southwest in 14 years, for an average of 4.1 years between commercially collectable crops. In local areas the interval between good nut crops may vary from as little as 2 to 5 years, or may be more than 10 years (Little 1941). Very rarely are good nut crops produced in consecutive years, such as occurred in 1991 and 1992 (Betancourt et al. 1993). It does appear that another good crop will occur in the fall of 1994, and if this occurs, then good crops will have been produced in 3 out of the past 4 years, a very rare occurrence. 193 Seed Yields Cones often occur on trees 3 to 4 feet in height and 10 to 20 years in age, and significant numbers of cones may be produced on trees that are 5 to 10 feet tall and 20 to 30 years old. However, the largest crops are produced on mature trees which are usually 20 to 30 feet tall with wide, full crowns. Pihons of this size may be 75 to 100 years old, and individual trees may produce cones for centuries (Phillips 1909, Botkin and Shires 1948, Ronco 1990). It should be noted that while seed production is indirectly related to tree age, seed production usually begins after a tree reaches a minimum specified size and height; for pinon that height appears to be 3 to 4 feet. Individual cones usually produce 10 to 20 seeds, but may produce up to 30 seed (Ronco 1990). During good seed years, good individual seed producers may produce up to 8 bushels of cones (Phillips 1909). At 3.3 pounds of seed per bushel and 1900 seed per pound (USDA Forest Service 1974), this translates to 26 pounds of seed or over 50,000 seed. In good seed years pinon stands produce an average of about 250 pounds of seed, or 475,000 seed, per acre (Kline 1993), but the best stands may produce as much as 300 pounds of seed per acre (Phillips 1909). In any particular year, seed crops are either good or poor, but seldom intermediate (Ronco 1990). The total pinon nut crop harvested annually in New Mexico, eastern Arizona, and southern Colorado averages between 1 and 2 mil- lion pounds. Crops harvested during good seed years probably average about 4 million pounds, with a range of 3 to 6 million pounds. The largest pinon nut harvest, which occurred in 1936, totaled almost 8 million pounds (Little 1941). The size of the nut crop harvested in good seed years is often regulated by the price that large seed dealers are willing to pay collectors; when seed supplies become plentiful, seed dealers lower the price paid to collectors. Increases in nut harvest during good seed years can be achieved simply by seed dealers providing modest increases in the average price paid to seed collectors from the cur- rent rates of $1-2 per pound to $2 or more per pound (Tanner and Grieser 1993). SEED PRODUCTION IMPROVEMENT Early seed production and seed yields per tree are under strong genetic control; consequently, seed production in stands can be increased substantially through individual tree selection. Seed production per acre can also be enhanced by concentrating seed production on the more productive pinon sites, increasing the number of good seed producers per acre, use of cultural treatments to increase seed production per tree, and by protecting the trees and stands from damaging insects and seed predators such as birds and rodents. By utilizing some or all of these methods and techniques, a relatively simple and cost efficient seed improvement program can be developed. Stand and Individual Tree Selection Currently, in the U.S., the pinon nut (primarily Pinus edulis Engelm.) is the only commercial nut crop collected entirely from "wild" trees (Little 1993). Some stands are inherently good and consis- tent seed producers and they are usually found on the more productive sites. These sites generally have deeper soils, with higher nutrient levels, and occur at higher elevations where annual precipita- tion is higher and better distributed throughout the growing season. A selection program can begin by conducting surveys on the more productive lands to locate the highest nut yielding stands. Identified stands should be marked and reserved for use as seed production stands. Over 50 years ago Little (1940) recognized that some individual pinon trees are consistently good nut producers, and others, consistently poor. He also noted that some trees produce more cones than others and some trees produce larger cones with more nuts per cone. The second phase of a selection program would be to identify individual high seed yielding trees within selected high seed yielding stands. The entire individual tree selection program could be accomplished within two to four years depending upon the number of individuals trees to be selected for inclusion in a seed production improvement program. Two types of individual tree selections can be made: Type 1 selections which are large, mature, high seed yielders that are to be left in place, and Type 2 selections which are relatively small (10-20 feet tall), young (15-20 years old), good seed producers, that can be transplanted into en- hanced seed production stands or into orchards. Enhanced seed production stands are those that have been selected as high seed yielders and enhanced by transplanting Type 2 selections among Type 1 selections, at fairly regular spacings to increase the number of high seed yielders per 194 acre. Seed production could be further enhanced in these stands by using one of more of the follow- ing cultural treatments: cultivation to remove competing vegetation, irrigation, fertilization, and shaping of the crowns to increase the cone bearing surface. SEED ORCHARDS The most intensive, and therefore most costly, way to increase seed production would be to es- tablish seed orchards on productive, agricultural type lands. Establishment and maintenance of pinon nut orchards would be similar to that of other nut orchards, such as the pecan and pistashio orchards in the Southwest. Two types of orchards could be established: transplant orchards or seedling seed orchards. Transplant orchards can be established by trans- planting Type 2 selections to an orchard site at regular spacings, such as 15 x 20 feet, after the site has been thoroughly prepared similar to an agricul- tural field. Seed orchards can also be established with seedlings produced from seed collected from Type 1 selections. This type of orchard may appear to have little promise since pinon typically exhibits very slow growth under natural conditions. How- ever, relatively fast growing seedlings can be pro- duced as containerized stock that are intensively cultured in a greenhouse-shadehouse growing complex. When grown under these conditions 6-8 inch tall seedlings can be grown in 6 months and 3-4 foot tall trees in 3 to 4 years. If intensively cultured, container-grown trees, are then planted to fairly productive sites where they can cultivated, irrigated, and fertilized, they can be grown to moderate seed production size in less than half the time it takes under natural conditions and 8-10 foot tall trees can be grown in 10 years from seed. During and after transplanting of Type 1 selec- tions or planting of intensively cultured seedlings, the orchards should be periodically cultivated, irrigated, and fertilized to promote seed produc- tion. Established orchard trees can be shaped to increase the cone producing surface of individual trees. In addition, a pruning program, designed to remove basal branches up to 5 feet above the ground, should be used to increase seed yields per cone, increased seed size, and increase yield of full seed per cone (Montano et al. 1980). The orchards should also be protected from cone and seed insects through maintenance of a intensively controlled pesticide application program and from seed predators, such as birds and rodents. Potential Yields from Orchards The following example of a small seed orchard illustrates potential seed production from a pinon transplant orchard. The type of orchard estab- lished, its size and age, and mix of cultural treat- ments used will determine actual amounts of seed that can be produced at any particular time after orchard establishment. Orchard: Type — transplant orchard Size — 10 acres Spacing — 15 x 20 feet Total trees— 1450 (145/acre) Assumptions: Time frame — 10 years after es- tablishment completed Average seed yield — 10 pounds per tree Good seed year every 4 years Seed Yields: 14,500 pounds of seed (1450 trees x 10 pounds/tree) every 4 years 3,600 pounds of seed per year 360 pounds of seed per acre per year Plus any seed produced in in- termittent years. A seed orchard of this size is probably not an economically sound venture because of the initial high unit costs of establishment and maintenance; however, yields estimated here can be easily con- verted for any multiples of 10 acres. Prior to any orchard establishment a series of economical analy- ses would have to be made in order to determine the minimal orchard size needed to provide a modest economic return from the orchard. SEED COLLECTION AND STORAGE Collection By far the vast majority of pinon nuts are picked by hand, one at a time, from the ground, after the majority of cones have opened. Perry (1922) estimated that 22 pounds of seed picked by hand is considered a fair day's gathering, although some especially dexterous persons can pick up to 195 40 pounds a day. More enterprising individuals spread sheets, blankets, tarps, plastic, etc. under individual trees, then shake the trees one or more times to dislodge seed from open cones. This method of collecting seed could be enhanced by mechanizing the collection process. This could be done by developing a harvester/shaker that would place a collecting surface around the base of the tree and thump or shake the tree to dislodge the seed. This method of collection would be most efficient if used in enhanced seed production stands or in seed orchards. In a seed orchard situation where the trees are all spaced at regular intervals, seed collection could be highly mecha- nized by use of a net retrieval system that is used in southeastern pine seed orchards (Edwards and McConnell 1982, McConnell and Edwards 1984). With this method of seed collection a net material, originally developed from carpet backing material, is laid in continuous strips under the trees prior to cone opening. When the majority of cones have opened naturally, individual trees can be shaken to dislodge seed. After essentially all of the seed has fallen onto the netting, the netting is retrieved by pulling individual netting strips from one end with a machine that collects the seed and places it into collection bins as it rolls the netting onto storage rolls. Seed Storage Routinely, unshelled pine seed can be stored successfully for up to 10 years, or more, when it is stored in sealed containers, at temperatures of 0 to 20 °F, with moisture contents between 5 and 10 percent. Unfortunately no published information is currently available listing the optimum, long-term storage conditions for pinon to maintain seed viability, and nutritional status and taste of the nuts. Pinon seed, however, can be stored under these conditions for a minimum of 10 years without appreciable loss in seed germination if the seed is of good quality when it is put into storage. Re- search and/or administrative studies are needed to determine the optimum long-term storage condi- tions necessary to maintain seed viability, and the high nutritional value and taste needed in the commercial nut industry. RECOMMENDED PINON SEED ENTERPRISE Currently, there are few commercial pinon seed dealers in the Southwest. Consequently, only relatively small amounts of pinon seed are avail- able on a consistent annual basis, and very little shelled seed is available to U.S. markets. However, there continues to be a very high demand for pi- non nuts as reflected by the volume of pine nuts imported annually into the U.S.. Delco et al. (1993) noted "that there may be a real window of oppor- tunity to market the pinon nut throughout the U.S. at this time" and "An understanding of the market place and the potential for developing a economic and environmental policy for growing, harvesting, shelling, and marketing pinon nuts is key to com- peting in the U.S. and world markets with these nuts". A group of individuals and/or an agency in the Southwest having extensive holdings of pinon lands could take advantage of this opportunity by establishing and operating a complete pinon seed enterprise which would include: 1) purchase of seed from local collectors, 2) establishment and maintenance of a seed improvement program to provide additional seed on a more regular basis, 3) long-term seed storage, 4) development and use of nut roasting and shelling equipment, 5) seed packaging, and 6) a seed marketing program. Development of this type of enterprise could result in a more stable supply of pinon seed at competi- tive prices for the food, landscaping, and reforesta- tion industries in the Southwest. LITERATURE CITED Barger, Roland L., and Peter F. Ffolliott. 1922. Physical characteristics and utilization of major woodland tree species in Arizona. USDA For. Serv. Res. Pap. RM-83, 80 p. Betancourt, Julio L., Elizabeth A. Pierson, Kate Aasen Rylander, James A Fairchild-Parks and Jeffrey S. Dean. 1993. Influence of history and climate on New Mexico pinon-juniper woodlands. In: Managing pirion-Juniper Ecosystems for Sustainability and Social Needs: Proc Symp. 1993, April 26-30, Santa Fe, NM, USDA For. Serv. Gen. Tech. Rep. RM-236: 42-62. Botkin, C. W, and L. B. Shires. 1948. The composition and values of pinon nuts. New Mexico Agric. Exp. Sta. Bull. 344, 14p. Delco, Steven, Roberta Beyer and Fritz Allen. 1993. U.S. market for imported pignoli nuts. In: Managing pinyon-Juniper Ecosystems for Sustainability and Social Needs: Proc. Symp. 1993, April 26-30, Santa Fe, NM, USDA For. Serv. Gen. Tech. Rep. RM-236: 164-167. Edwards, Jerry L., and James L. McConnell. 1982. Forest tree seed harvesting system for loblolly pine. ASAE Paper No. 82.1589, Winter Meeting, lOp. 196 Eremko, R. D., D. G. W. Edwards and D. Wallinger. 1989. A guide to collecting cones of British Columbia coni- fers. Joint Publ., Forestry Canada and the British Co- lumbia Ministry of Forests, FRDA Rep. 055, 114p. Kline, Jeff. 1993. My vision of the pinyon/socioeconomic potential of pinyon woodlands. In: Managing pinyon-Juniper Ecosystems for Sustainability and Social Needs: Proc. Symp. 1993, April 26-30, Santa Fe, NM, USDA For. Serv. Gen. Tech. Rep. RM-236:3-8. Lanner, Ronald M. 1981. The pinyon: A Natural and Cultural History. Univ. Nevada Press, Reno, NV, 208p. Little, Elbert L., Jr. 1938a. The earliest stages of pinyon cones. USDA For. Serv., Southwest. For. and Rng. Exp. Sta. Res. Note 46, Tucson, AZ, 4p. Little, Elbert L., Jr. 1938b. Stages of growth of pinyons in 1938. USDA For. Serv, Southwest. For. and Rng. Exp. Sta. Res. Note 50, Tucson, AZ, 4p. Little, Elbert L., Jr. 1940. Suggestions for selection cutting of pinyon trees. Southwest. For. and Rng. Exp. Sta. Res. Note 90, Tucson, AZ, 3p. Little, Elbert L., Jr. 1941. Managing woodlands for pinyon nuts. Chron. Bot. 6(15):348-349. Little, Elbert L., Jr. 1993. Managing southwestern pinyon-juniper woodlands: the past half century and the future. In: Managing pinyon-Juniper Ecosystems for Sustainability and Social Needs: Proc. Symp. 1993, April 26-30, Santa Fe, NM, USDA For. Serv. Gen. Tech. Rep. RM-236: 105-107. McConnell, James L., and Jerry L. Edwards. 1984. The net retrieval seed collection system for Southern Re- gion seed orchards — an economic study. In: Proc. Third Biennial Southern Silicultural Research Confer- ence 1984, Nov. 7-8, Atlanta, GA, USDA For. Serv. Gen. Tech. Rep. SO-54: 252-254. Montano, J. M., J. T Fisher and R. E. Gomez. 1980. In- crease of pinyon nut size by basal pruning. Hort. Sci. 15(6):727-728. Perry, Walter J. 1922. A word for the lowly pinyon. J. For. 20:521-526. Phillips, F. J. 1909. A study of pinyon. Bot. Gaz. 48:216-223. Ronco, Frank E, Jr. Pinus eduli Engelm. pinyon. In: Burns, Russell M., and Barbara H. Honkala, tech. co- ord. Silvics of North America, Vol. 1 Conifers, USDA For. Serv, Agric. Handb. 654:327-337. Tanner, Ellis, and Don Grieser. 1993. Four generations trading pinyon nuts with native Americans: changes needed for future prosperity. In: Managing pinyon-Juniper Ecosystems for Sustainability and Social Needs: Proc. Symp. 1993, April 26-30, Santa Fe, NM, USDA For. Serv. Gen. Tech. Rep. RM-236:29-33. U.S. Department of Agriculture, Forest Service. 1974. Seeds of woody plants in the United States., C. S. Schopmeyer, tech. coord., USDA, Agric. Handb. 450. Washington, DC, 883p. 197 Carrizo Demonstration Area Restoration of a Southwest Forest Ecosystem Richard S. Edwards1 Abstract. — The Carrizo Demonstration Area is located on the Smokey Bear Ranger District, Lincoln National Forest. It encompasses 55,000 acres of National Forest and private land, and is comprised mainly of pinon-juniper forest. The Carrizo Demonstration Area was established in 1989 as a pilot project designed to restore and sustain watersheds, increase natural food production for wildlife and livestock, and increase biological diversity by managing the area based on ecological principles. The Carrizo program is a demonstration of the Forest Service's ecologically based, multiple re- source management. The primary purpose is to provide stewardship of the land to achieve and sustain desired conditions, cooperative partnerships to plan and implement projects, and utilize research and technology to pro- vide quality on-the-ground resource management and protection. Over 4,000 acres of multi-resource improvements have been planned and im- plemented thus far to begin a transformation of the area toward desired fu- ture condition. The desired future condition will be achieved when active accelerated soil erosion is stopped, steep gully slopes are stabilized, and permanent riparian vegetation is restored. A mosaic of vegetative structural age classes and densities will exist within the different ecotypes, moving toward a balanced and stable ecosystem. Enduring partnerships with land- owners and permittees will be permanently established to aid in sustaining the desired condition of the land. INTRODUCTION The Southwestern Region is gearing up to im- plement a program that emphasizes an ecological approach to multiple-use management of the pinon-juniper ecosystem. The Lincoln National Forest started on a project about four years ago, called the Carrizo Demonstration Area. Much of the Carrizo area now contains large expanses of continuous canopy pinon-juniper forest. Under these present conditions, natural openings are dominated by young pinon (Finus edulis) and juni- per (Juniperus monosperma, Juniperus deppeana, and Juniperus scopulorum) trees, and historically open woodlands have become dense thickets. Due to the increased competition from trees, these ecosystems are devoid of the grasses and other vegetation that hold the soil in place, contribute to plant diversity, and provide food or cover for various wildlife and 1 Forester, USD A Forest Service, Lincoln National Forest, Smokey Bear Ranger District, Ruidoso, NM. livestock. Much of the productive soil beneath these dense woodland stands has eroded away, leaving behind an extensive gully system which continues to transport silt-laden water into streams and rivers, and serves to lower the water table. The need for this project was brought about through the urging of area private landowners and grazing permittees, who for years have had to contend with the deposition of millions of tons of sediment that originated on National Forest land, as well as a steady decline in livestock grazing capacity due to a decrease in forage. The pinon-juniper woodlands have gone through many changes over the past 20,000 years. Due to gradual global warming, they have migrated from lower elevations to higher elevations and extended their range from southern latitudes to northern latitudes (Betancourt et al. 1986). By the middle of the 19th century, most of the 198 pifion-juniper woodlands in south-central New Mexico were located on steeper, rockier slopes, although transition zones existed between pirion- juniper and the short-grass rangelands and pifion- juniper and ponderosa pine (Pinus ponderosa). Much of the woodland area, in particular the lower elevation zone, was very open in appearance. It had been maintained in that condition by periodic fire. Tree ring studies in New Mexico indicate that many forests burned, on the average, at 7 to 10 year intervals (Stoddart et al. 1975) prior to settlement of the area. But one of the most remarkable changes oc- curred during the last 100 years. During the late 1800's and early 1900's, much of New Mexico re- ceived intense grazing pressure from domestic livestock. Lincoln National Forest records show that by 1902 on what is now the Smokey Bear Ranger District, 80,000 head of livestock were grazing on the Lincoln Forest Reserve (Hightower, 1902). To put this in perspective, today the permit- ted livestock - the grazing capacity is 5,000 head of livestock. Because of this heavy livestock grazing, the grasses were reduced to the point where they could no longer carry these periodic wildfires which kept the pinon and juniper trees in check. The reduction in available fuels in combination with fire suppression by public land management agencies, led to a proliferation of young pinon and juniper trees that throughout this century have increased, and are out-competing native grasses and forbs. As tree canopies became closed, grasses and plants that prevented erosion and provided forage for wildlife and livestock rapidly declined because they could not compete with the pinon and juniper trees (Evans, et al. 1988). Many of the perennial streams and springs, life-blood for the rich diversity of riparian and wetland ecosystems, were in part lost because of the excessive water requirements of these woodland trees (Ponce and Lindquist, 1990). Livestock producers throughout the Southwest have for many years been concerned because of the long-term loss of forage productivity associated with this situation. Deer, elk, wild turkey, many songbirds, and other species of wildlife have been adversely affected by this change in habitat conditions. Private landowners adjacent to the National Forest have had to contend with the deposition of millions of tons of sediment that originated on the forest. Water, a scarce and precious commodity throughout the southwest, requires healthy forest watersheds. The quality, and potentially the quantity of water supplies for nearby communities, agricultural centers in Pecos River Valley and Tularosa Basin, and local wildlife and livestock are affected by the condition of the watershed. MISSION Our mission for the Carrizo area is to establish cooperative partnerships to aid in the development of sound land stewardship principles and to serve as examples in the implementation of land man- agement activities to restore watersheds to satisfac- tory condition. Stewardship goals also include providing for a variety of wildlife habitats, increas- ing plant and animal diversity, restoring the natu- ral beauty of the landscape, and improving overall ecosystem health. Management strategies focus on soil stabilization practices, vegetation management, water resource development, vehicular travel management, and sound range management prac- tices, and are based on the best scientific and man- agement information available. The desired future condition will be achieved when active accelerated soil erosion is stopped, steep gully slopes are stabilized, and permanent riparian vegetation is restored. A mosaic of vegeta- tive structural age classes and densities will exist within the different ecotypes, moving toward the balance and stability which occurred prior to Euro- pean man's settlement of the area. Prescribed fire will be introduced to resemble the natural fire frequency that evolved with, and shaped the natu- ral ecosystem. Enduring partnerships with adja- cent landowners, traditional and non-traditional users, and nearby communities will be perma- nently established to aid in sustaining the desired condition of the land. ECOLOGICAL APPROACH The focus in the development of the Carrizo project was the recognition that all resources are interrelated and the integration of all resources into a management system is essential for long term success. Each aspect of the project was evalu- ated for it's effects on all resources, including the human environment. Our past custodial manage- ment philosophy for pinon-juniper ecosystems has led to a steady decrease of resource values (Doughty, 1987). The Carrizo area interdisciplinary planning team devised strategies to restore and sustain woodland watersheds. The major identified 199 CARRIZO DEMONSTRATION AREA TERRESTRIAL ECOSYSTEM SURVEY MAPPING UNIT PRESCRIPTIONS TES UNIT TOTAL ACRES SLOPE DFC DESCRIPTION MAX. OPENING SIZE 3 6,284 U " IQ/O vjiaoolailu/Oavaririari 10-200 Ac. 7 2,853 n - 1 ^% \J \ sj /O oavai ii leu 1/ 1 o vvuvjuiaiiu 1 - 30 Ac. 8 3,908 15- 40% PJ Woodland/Oak Woodland 1 - 10 Ac. 11 1,404 w 1 w /O vji ciooicii iu/ oqvhi i i mi i 10- 100 Ac. 265 1,449 0-1 5% V 1 >J /o .Qfl\/flnnflh/P. I WrwvJtanH Ouv cii 1 1 1 qi 1/ r o v v icii iu 1 - 30 Ac. 3014 1,499 + 40% Oak Woodland S-/ C4 rv TV W W U lull v3 0 - 4 Ac. 302 3,154 1 5 - 40% PonHpro^fl/P. 1 Woodlflnd 0 - 4 Ac. 3034 5,057 + 40% PonHprn^fl/P.I Woodland 0 - 4 Ac. 3054 2,212 4- 40% Mivpd f^onifoT 1 VII ACU VUI 1 1 ICI 0 - 4 Ac. 3074 2,209 + 40% Mixed Conifer ITIIAVM ***r\S1 III wl 0 - 4 Ac. 311 3,527 0-15% Ponderosa/PJ Woodland 0- 10 Ac. 336 2,516 15-40% Savannah/PJ Woodland 1 - 20 Ac. 3404 1,276 + 40% Oak Woodland 0 - 5 Ac. 3484 2,094 + 40% Oak/Mt. Mahogany Open 401 2,179 0-15% Savannah/PJ Woodland 1 - 30 Ac. 402 5,781 15-40% PJ Woodland/Oak Woodland 1 - 1 5 Ac. 404 1,345 0 - 1 5% Ponderosa/PJ Woodland 0 - 1 0 Ac. 405 2,475 1 5 - 40% Ponderosa/PJ Woodland 0- 10 Ac. elements of this program are watershed, wildlife, vegetation, ecology, and range management. Through the use of the Southwestern Region's Terrestrial Ecosystem Survey the team identified high priority potential treatment areas as those with unsatisfactory watershed condition and high soil productivity. The Terrestrial Ecosystem Survey is also used as the basic ecological unit to display objectives and prescriptions for desired future condition. The planning team compared the exist- ing condition with the desired condition to develop a list of possible management practices and pre- scriptions for each Terrestrial Ecosystem Survey mapping units. The following table displays the major Terres- trial Ecosystem Survey mapping units and their desired condition descriptions: With the help of cooperative partnerships, treatments to produce desired conditions have included rehabilitating gullies by constructing small dams and reshaping gullies; establishing native vegetation to stabilize the soil by thinning trees for fuelwood, removing unwanted trees excess trees through mechanical means, pre- scribed fire, and by reseeding disturbed areas; providing dependable water supplies for wildlife by restoring and protecting riparian areas, in- stalling inverted umbrella trick tanks, and devel- oping existing springs; increasing overall forest health through thinning or harvesting trees in diseased or overstocked timber stands and through prescribed fire; and establishing travel access in line with resource needs by closing or obliterating unnecessary roads, relocating roads to more stable or suitable areas, and maintaining necessary roads and trails. STEWARDSHIP The restoration of watersheds is designed to stop excessive downstream sedimentation, pre- serve soil productivity and increase the duration of channel flows. In addition to stabilizing water- sheds, benefits of the ecosystem approach being implemented include increased wildlife habitat capability, improved rangeland condition, in- creased visual diversity, and an increase in supply of forest products such as fuelwood, fence posts, vigas and poles. Once ecological restoration is established, the emphasis will be on sustaining a healthy ecosystem. Sound range management practices, such as deferred rotation grazing, fuel- wood harvest, and the use of prescribed fire to maintain diversity, will be used to achieve a sus- tainable ecosystem. Where treatments have been implemented, watershed conditions have im- proved dramatically. Cool season native species of grass and forbs which were once thought to be lost have returned in abundance. In several drainages, springs have begun to flow again, creating many opportunities to establish or enhance riparian vegetation. As a result of these changes, many species of wildlife which were declining in num- bers have returned to the area. A more diverse setting across the landscape has increased the scenic quality of the area, and will allow future resource management to more easily maintain a natural appearance. Positive changes have even begun to occur on adjacent private land following treatments ac- complished on National Forest. In one area, a pond located on private land had filled with sediment from past gully and sheet erosion transported by 200 overland flow from the National Forest. The land- owner removed 4,800 cubic yards of topsoil from this pond at the same time watershed restoration treatments were being implemented above the pond on the National Forest at the same time wa- tershed restoration and vegetation treatments were being implemented above the pond on the Na- tional Forest. During Spring season, a spring which had not run in at least 35 years began to flow. The large spring, as well as many other new, but smaller springs in adjacent drainages, continued to flow throughout the summer, filling the pond with clean, clear water. In addition to baseflow increases, sediment coming from the National Forest was minimal. The landowner was able to stock the pond with trout and catfish, and is now the per- manent summer residence for many waterfowl. Opportunities to improve economies within the surrounding rural communities have been en- hanced due to increased production of forest products such as fuelwood and poles for vigas, and an increase in big game wildlife. Partnerships with adjacent landowners and others have opened up new lines of communication and have substantially increased the level of trust with our public. PARTNERSHIPS WITH PEOPLE Partnerships are an integral part of this effort. Thirteen grazing permittees, three adjacent private landowners, New Mexico Department of Game and Fish, New Mexico Division of Forestry and Resource Conservation, New Mexico State Univer- sity (NMSU), New Mexico Range Improvement Task Force, and NMSU Cooperative Extension Service participated in long range project devel- opment. Numerous field trips involving diverse groups of constituents have been hosted to inform the public of the need for a stewardship approach to management of the pinon-juniper ecosystem. Congressional representatives have been closely involved throughout both the planning and initial implementation phases of the project. Grazing permittees and private landowners have been the primary partners with the Forest Service for site specific watershed restoration and vegetation management projects. Project implementation partnerships are designed to meet multi-resource objectives by achieving complete treatments. For example, commercial fuelwood cutters have historically harvested only those trees which can be sold for firewood, leaving hundreds of excess trees per acre. Commercial fuelwood cutters within the Carrizo area now cut all trees not designated to be left. In fact, some partners can harvest a fuelwood area by written prescription, no longer needing the Forest Service to designated leave trees, creating additional savings to the gov- ernment. Private landowners have purchased fuelwood sales on the national forest, and perform the same treatment on their adjacent private land. One landowner even entered into a cooperative agreement where vegetation on both National Forest and private land was managed with a pre- scribed burn. In Fiscal Year 1993, a partnership involving South Central Mountains Rural conservation and Development, the Administrative Council of the Western Region's Sustainable Agriculture Research and Education Programs at the University of Cali- fornia at Berkeley, New Mexico State University and the Forest Service was formed to produce and distribute a high quality video portraying ecosys- tem management with the Carrizo Demonstration Area. The primary objective of the video is to edu- cate a wide range of publics and develop support for an ecological approach to multiple-use man- agement in the pinon-juniper woodlands. Partici- pants in the video include agencies and environ- mental groups such as Soil Conservation Service, Forest Service, New Mexico State University, Na- ture Conservancy, American Wildlands, as well as many private individuals. COLLABORATION WITH RESEARCH An ongoing focus of the Carrizo project has been to attract interest from researchers to explore the many questions associated with managing pinon-juniper woodlands on a landscape scale. Many institutions, organizations and individuals are involved in ongoing pinon-juniper research. The Rocky Mountain Forest and Range Experiment Station is currently researching on-site soil pro- ductivity and modeling soil erosion in the Carrizo area. Other efforts include Southwestern Region's New Mexico pinon-juniper management initiative, U.S. Department of Agriculture's cooperative Pecos River Basin Study, NMSU Cooperative Extension Service's rangeland watershed program, and New Mexico Department of Game and Fish - Habitat Improvement Stamp (Sikes Act) program. 201 PROJECT REVIEW Since the inception of the Carrizo Demonstra- tion Area in 1989, a number of projects have been implemented which are moving the area closer to the desired condition. With the help of cooperative partnerships, approximately 2,500 acres of unsatis- factory condition watershed have been treated through vegetation management to increase herba- ceous ground cover, four miles of gullies have been treated through installation of structural improve- ments or gully sideslope stabilization, and five miles of roads have been obliterated to reduce another source of downstream sedimentation. Specific improvements for wildlife habitat have been implemented on almost 1,100 acres through prescribed fire or creation of wildlife openings. In addition, two wildlife water developments were installed, and 15 acres of existing riparian have been fenced to manage livestock grazing. Forest products sold as a result of vegetation treatments include 2,850 cords of fuelwood, 4,000 board feet of timber, and 500 small and medium poles. Implementation of another project is underway to improve habitat for big game wildlife, as well as northern goshawk, using prescribed fire. Except under extreme conditions, use of prescribed fire to create openings within most areas of pinon-juniper is very difficult to accomplish successfully. This project was designed to thin seedlings and saplings prior to burning to create the necessary ground fuels to carry the fire. The fire will then result in a natural appearing mosaic of different habitat struc- tural stages across the landscape. The natural food supply for big game wildlife as well as goshawk will be increased, and watershed conditions will be improved through increased ground cover. CONCLUSION The management situation in northern New Mexico is different than in south-central New Mexico, primarily from the cultural value stand- point and usage of pinon-juniper woodlands. But the ecological condition is essentially the same. Watersheds are being severely degraded to the point where site productivity is being lost. We cannot afford to lose much more topsoil from our woodland watersheds without seriously endanger- ing production of commodities such as the pihon nut crop and fuelwood. As pointed out earlier, we have already experienced the loss of understory vegetation critical to wildlife and livestock And possibly worst of all, damage to riparian areas has been extensive. The ecological approach to multiple use man- agement is a win-win proposition. Take pinon nut production for instance. Years of research and implementation have shown that if you thin se- lected pinon-juniper sites, larger pinon trees can be grown, and thereby increase the production of pinon nuts. By lopping and scattering slash from the tlunning, ground cover would be increased, reducing erosion. And as shown by projects im- plemented within the Carrizo area, diversity for all resources would be increased. For Carrizo, one of our basic objectives is to test different treatments for managing woodland wa- tersheds. Some of the treatments will not respond the way they are designed. But the point is, we have already learned a great deal from past mis- takes and successes, and we will continue to moni- tor our projects to learn and make the necessary adjustments to achieve the desired future condi- tion. The Southwestern Region of the Forest Serv- ice has already taken a major step forward in rec- ognizing the values and complexities of the pinon- juniper ecosystem. REFERENCES Betancourt, Julio L. 1987. Paleoecology of Pinyon-Juniper Woodlands: Summary. Proceedings - Pinyon-Juniper Conference. USDA Forest Service General Technical Report INT-215: 129-139 Doughty, Jim W. 1987. Problems With Custodial Man- agement. Proceedings - Pinyon-Juniper Conference. USDA Forest Service General Technical Report INT- 215: 29-33 Evans, Raymond A. 1988. Management of Pinyon- Juniper Woodlands. USDA Forest Service General Technical Report INT-249. High tower, Clement 1902. Grazing Report from Forest Supervisor, Lincoln Forest Reserve. Ponce, Victor M. and Lindquist, Donna S. 1990. Man- agement of Baseflow Augmentation: A Review. Water Resources Bulletin - American Water Resources As- sociation: 259-268 Stoddart, L.A., Smith, A.D., and Box, T.W. 1975. Range Management Third Edition. McGraw Hill Book Co. St. Louis, MO. 202 Silvicultural Systems for Pinon- Juniper James R. Ellenwood1 Abstract. — Silvicultural systems and cutting methods can be applied to pinon-juniper stands. Several silvicultural systems are available depending upon the objectives of the landowner. Even-aged, uneven-aged, and irregular-aged silvicultural systems are described and compared for managing pinon-juniper stands. Each system involves the application of treatments to individual stands to achieve a desired condition. Silvicultural treatment an be maintained. A proposed density management regime is presented. INTRODUCTION The pinon-juniper woodland is the largest cover type in the southwestern United States cov- ering 47 million acres of land, however, despite this superlative, it has often been considered the least managed. Low demand and product value for pinon-juniper motivated past management to emphasize forage values, however, the suppression of wildfire and periods of over-grazing has conse- quently reduced forage production. Over the last two decades, demand for fuelwood has increased dramatically, which has resulted in an increase in demand for fiber products from the pinon-juniper ecosystem. In more recent years there has been a paradigm shift towards ecosystem management, as a result, two major initiatives have been adopted by the USFS, the Pifion-juniper Initiative and the Forest Health Initiative. The intent of these initia- tives is to restore the health of the pinon-juniper ecosystem. Silvicultural practices are some of the tools that can be used to accomplish this objective. The intent of this paper is to present some of these available tools and to present them within the perspective of ecosystem management. SILVICULTURE AND ECOSYSTEM MANAGEMENT Silviculture is the most ancient conscious application of the science of ecology; the association arose before the word "ecology" was coined (Smith 1986). Silviculture relies upon ecological knowledge out of necessity. While ecosystem management is perceived to be a philosophical shift in management emphasis, the fundamentals of silviculture remain profoundly rooted in ecological principles. In other words, silviculture uses an ecological approach to accomplish management objectives. The basic management unit for silviculture is referred to as a stand or site. A site is essentially the forest community from which tree life forms have a consistency of age, density, species, and structural composition. Silvicultural systems are applied on a site basis. The aggregate of many sites over a large area (10,000 to 100,000 acres) is consid- ered a landscape. Depending upon landscape objectives, several silvicultural systems can be applied to various sites throughout a given land- scape. Ecosystem management emphasizes the bigger picture by looking at units of land over varying scales of time and space. An ecosystem manage- ment approach develops objectives for the land- scape, from which the site objectives are derived to accomplish these landscape objectives. A specific site objective is not necessarily the "best" one for a given site, but can be one that contributes to the overall landscape objectives. The aggregate of many site objectives will meet the landscape objec- tives. For example, at any given time, there is a portion of the landscape providing forage, another portion providing old growth characteristics, and another portion providing wildlife cover. Thus, silvicultural practices are used to main- tain or establish forest ecosystems comprised of a 1USDA Forest Service, Kaibab National Forest, Williams, AZ. 203 changing mosaic of successional stages and vege- tative patterns. Variations in individual sites can range from highly diverse forest stands of multiple species and structures to areas of relatively uniform vegetation and from areas of openings to areas of unfragmented continuous tree canopies (Bergsvik 1994). ECOLOGICAL POTENTIAL, TES, AND THE APPLICATION OF SILVICULTURE While the pifion-juniper woodland is often viewed as a single cover type, there is great diver- sity in its classification. Plant associations are used to describe the potential natural community. The potential natural community is often described in ecological surveys to determine site potential. Plant associations are used to determine what species composition and productivity can be man- aged on a site. By characterizing the potential of a site, comparisons can be made between various management strategies, for example, a site that has a strong woody vegetation potential may be too costly to maintain in a grassy condition. The Terrestrial Ecosystem Survey uses the plant association and considers climate and soils. It is estimated that approximately 280 map units will eventually be described for the pifion-juniper cover type in the Southwestern Region. The Terrestrial Ecosystem Survey has been used to segregate landscapes into areas of common potential. From this, site specific data is collected and an existing condition is determined. A site specific application of silviculture is documented as a silvicultural prescription. Silvi- cultural prescriptions are developed for a given site based upon the existing condition, the ecological potential of a site, the silvical characteristics of the desired species, and the management objectives as defined by the desired condition. DESIRED CONDITION Desired conditions are selected based upon management objectives. The desired condition needs to be cognizant of the ecological requirements of each of the desired species. Potential natural community is not necessarily the desired condition, but one of the bounds from which one is derived. For example, it may be desirable for a preferred wildlife species to maintain a serai successional stage by maintaining cool season grasses in a climax pifion-juniper plant association — the plant species desired need to be ones that are capable of growing on the site. The existing vegetation condition may also not be desired, for example, a site in which grasses are absent in the interspaces of pinon-juniper groups, the desired condition would be to have grasses in the interspaces. There are three generalized management ob- jectives which are often applied to pifion-juniper management: Sustaining grassland - on areas which have historically been grasslands that have slowly re- verted to pinon-juniper. This objective has been the subject of much debate. Since the principle objective is to maintain grassland species, the treatments applied are not considered part of a silvicultural system. However, some of the silvicul- tural treatments can be used to achieve this objec- tive. Sustaining woodland - on areas which have historically been pinon-juniper. This objective relies on the application of traditional silvicultural systems. Sustaining woodland does not preclude the production of forage, since this creates a mosaic of stand structures throughout the landscape and each site and portions of sites periodically provides forage. Sustaining woodland savannah - on areas where the historic density of pifion-juniper was relatively low and the grasses dominate the under- story. This objective uses a modified application of silviculture since the principle objective is not purely managing the woodland, but managing the matrix of grassland and woodland upon the same site. One needs to account for the regeneration of the savannah woodlands the same as one needs to for the woodland objective. Maximizing a single resource such as fiber or forage over a landscape may no longer be the desired condition under ecosystem management. A healthy ecosystem can provide some of those products and sustain the many parts that com- prise it. Implementing a silvicultural prescrip- tion does not mean to maximize fiber produc- tion. Based upon landscape objectives, a desired condition is selected for each site. It is through the desired condition that site objectives are set. The practice of silviculture offers a means to manipulate vegetation to meet a variety of man- agement objectives. These practices are not the vehicle for the establishment of the objectives, they are simply the means of achieving them (Bergsvik 1994). 204 SILVICULTURAL SYSTEMS Silvicultural systems for pirion-juniper wood- lands have been previously described by Bassett (1987), Ronco (1987), and HoUand (1989). However, given the new directions of ecosystem manage- ment, a fresh look needs to be presented. As in most sciences, terminology is often con- fusing. There is a distinction between silvicultural systems, methods, and treatments, which are often misapplied. A silvicultural system consists of a series of treatments that regenerate and control the stocking of a site through time. A regeneration method consists of one or more treatments used to establish a new forest on a site. A silvicultural treatment is usually a single operation applied to site. There are 2 broad classes of silvicultural sys- tems, high-forest systems, which rely upon regen- eration by seeds, and low-forest systems (coppice), which rely upon regeneration by vegetative propa- gation such as sprouting. The low-forest system is dependent upon the ability of a desired species to vegetatively propa- gate. In the pinon-juniper woodland, only one species, alligator juniper, is considered a prolific sprouter (Barger and Ffolliott 1972). This system is typically included under the clearcut method of the high-forest system, however, it is more correct to recognize the coppice method as its own distinct system based upon the regeneration source. If the desired condition is to maintain alligator juniper in a pure stand, then the coppice method is an option to consider. In Region 3, the high-forest system is sub- divided into three systems: even-aged, uneven- aged, and irregular-aged systems. Even-aged Systems Even-aged systems rely primarily on maintain- ing trees within a site at or near the same age with variations of no more than 20 percent of the site rotation. Even-aged systems utilize separate re- generation and tending treatments. Even-aged regeneration can be accomplished with three methods. Clearcut Method - removes all trees from a given site in order to allow the stand to regenerate. Regeneration is achieved from seeding by adjacent areas, seed stored in the litter of the prior stand, or by planting. This treatment has the same appear- ance as a grassland restoration treatment, however the difference lies in the long term objective, which is to regenerate trees rather than to maintain grassland. The clearcut method is the simplest method to implement as well as the least expensive if planting can be avoided. This is the most effec- tive method to temporarily increase forage and browse. Woodland species produce large, wingless seeds which do not disperse very well by wind, however, bird species such as the corvids, have been well documented to disperse pinon and juniper seeds (Balda 1987). Typically, seeds are cached in areas that remain accessible during the winter, such as in tree cavities and the base of large trees. It is generally accepted that the clearcut method is less than reliable for establishing a prop- erly stocked stand, however, the presence of cached seeds and presence of advanced reproduction lessens this problem. The clearcut method is also effective at controlling the larger dwarf-mistletoe infected areas. Seed Tree Method - removes most of the trees from a given site leaving behind a few seed trees which provide the seed source for the next stand. This treatment has the same appearance as a woodland savannah treatment, however the dif- ference lies in the long-term objective, which is to regenerate the site rather than to maintain the woodland savannah. This method is subject to the same dispersal limitations of the clearcut method. Shelter wood Method - opens the crown can- opy enough to encourage seed production and reduces competition for light and moisture to allow for the establishment of new seedlings, while leaving enough trees to protect the site from the drying effects of wind and too much light. This method is effective in controlling smaller dwarf- mistletoe infected areas. This method is usually implemented in several cuts or steps over a num- ber of years. The prepatory cut is used to develop windfirm- ness in leave trees, crown development for seed production, and accelerate the breakdown of a deep duff layer. None of these objectives appear to be necessary for most pinon-juniper stands (Bassett 1987). The shelterwood seed cut leaves the best seed producing trees to establish seedlings. This treat- ment can coincide with the objective of providing pinon nut production since it requires the selection of good seed producing trees to provide for the next generation of trees. The removal cut removes the sheltering seed trees to allow the newly established seedlings and saplings to grow freely This treatment may be performed in one or more cuts. The period of time 205 between the seed cut and removal cut may be as long as 40 years for a stand with a 200-year rotation age or twenty percent of the established rotation age of the stand (Bassett 1987). Under the final removal cut, one has the option of leaving reserve trees for continued production of pirion nuts or future snags. When an adequately stocked understory of ad- vanced reproduction exists, a simulated shelter- wood method can be applied (Ronco 1987). The seed cut is not necessary and removal cuts are performed to release the established seedlings and saplings. Pinon seedlings and saplings are typi- cally found underneath the canopy of the mature pinon. Caution should be taken in that the re- moval of the overstory cover may kill these seed- lings. Mortality is reduced when seedlings are greater than one foot in height (Gotfried 1993). Each regeneration treatment may include as- sociated treatments. Prescribed burning is some- times used to reduce slash and control competing vegetation as a preparation for the seedbed. Intermediate Treatments - are applied to a site between regeneration events. Sanitation-salvage treatments are used to remove dead or dying trees, often to control an insect or disease outbreak. Pre- commercial tWnning is used to control the stocking of young trees which are too small for commercial value. Commercial thinning is used to control the stocking of larger trees which have a commercial value. However, this is rarely practiced in the pifion-juniper woodland because low product value and high operating cost makes profitability highly marginal. Uneven-aged Systems Uneven-aged systems involve the manipulation of a site to simultaneously maintain continuous forest cover, regenerate desirable species, and to allow for growth and development of trees through a range of age classes. Within a given stand, the regeneration, tending, and harvesting treatments occur simultaneously. Uneven-aged systems require setting up a residual stocking level, a maximum diameter, a desired diameter distribu- tion (usually described by a factor of q), and a cutting cycle. Because of the slow growth of woodland species, a cutting cycle of 50 to 100 years is recommended (Meeuwig 1983). A 40 year cut- ting cycle may be appropriate for woodlands stands which are capable of producing 10 or more cubic feet of wood per year (Bassett 1987). There are 2 methods of uneven-aged systems. Single-tree selection method - selects individ- ual trees in various diameter classes for removal. This method works well with shade tolerant spe- cies. While woodland species are considered shade intolerant, seedlings survive in moderate shade, often growing underneath the protection of mature trees and shrubs (Meeuwig 1983). This method is considered the best method to fit most natural pinon-juniper stand conditions. Regeneration is more reliable because the heavy seeds do not have to travel far to adequately occupy the smaller re- generation units, in addition, shading from the residual trees helps to protect the established seed- lings (Bassett 1987). This method is effective for controlling dwarf-mistletoe infections which occur on a few isolated trees. Single-tree selection is a method to consider if its desired to re-establish the pifion-juniper and grass matrix where small inter- spaces between groups of trees are occupied by a grass component. Group selection method - selects trees in groups throughout a stand for treatment. Regen- eration openings are recommended to be no larger than twice the average height of the mature trees in the site (Bassett 1987). This method works well with shade intolerant species and can be used in pinon-juniper woodlands. This method is also limited by the previously stated dispersal limita- tions. This method is effective in controlling dwarf-mistletoe infections which occur in groups of trees. Group selection is another method to con- sider if its desired to re-establish the pinon-juniper and grass matrix where larger interspaces between groups of trees are occupied by a grass component. Irregular-aged Systems Irregular-aged systems are not recognized as standard silvicultural systems, but are a combina- tion of the other two systems. This system relies upon maintaining trees in a multi-storied condition much like the uneven-aged system, however, not all ages are present. This system works well where regeneration is not a consistent periodic event and where the desired stand condition is not purely even-aged or uneven-aged. The group shelter- wood method has been applied in the southwest to partially mimic the grouped distribution of pon- derosa pine. This results in a stand that is multi- aged, but not all-aged. The application of this system to pinon-juniper woodlands has not been previously discussed and is being offerecf here as an additional tool for consideration. 206 Each of these practices can be applied at differ- ent levels of intensity to meet site specific objec- tives that address local conditions. SILVICULTURAL PRACTICES AND RESEARCH Pirion-juniper has been managed with many treatments applied throughout the last century Many of the treatments have concentrated on restoring the grassland cover type. Given the vast amount of area, only a small portion of area has had a silvicultural system implemented. To illus- trate the application of silvicultural practices three project areas are described. Heber Ranger District Study Plots - a few silvi- cultural prescriptions are being studied by the RMFRES as part of an on going research project. Three treatment types are being studied, grassland conversion, simulated shelterwood removal cut, and individual tree selection. The single tree se- lection is based upon a Q of 1.2 (computed on a 1 inch diameter class), a maximum diameter of 15 inches with the upper limit being flexible depend- ing upon the available trees, and a residual basal area of approximately 60 square feet (Gotfried, pers. com.). This study is currently on-going and the results will not be determined for several years. In general, the District sees this as an opportunity to implement more projects in the pinon-juniper. Carrizo Ecosystem Management Project on the Lincoln National Forest has implemented several prescriptions for ecosystem management over the last 4 years. The primary objective for the project is for restoring the watershed condition. Several fuelwood treatments and pushes have been ap- plied to restore what was previously a pinon- juniper savannah. Several thinning treatments have been applied to generate scattered slash in the interspaces. This provides shelter to protect the site and to establish grasses. Group shelterwood treatments were applied to sustain the pinon- juniper woodland (Edwards, pers. com.). BIA-Albuquerque has set up several study ar- eas on tribal lands throughout New Mexico to examine the effectiveness of several silvicultural treatments, such as single tree selection, group selection, shelterwood seed cuts, and pinon nut production thinnings. In addition to the study areas, the BIA actively manages the pinon-juniper woodland for fuelwood and pinon nut production. Many of the treatments are focused upon sanita- tion/salvage and clearcutting to control dwarf- mistletoe and Ips beetle outbreaks. Pinon nut Table 1. — Single-tree Selection Diameter Distribution for Pinon-juniper. Diameter class TPA BA SDI 0.1 to 3.4" 130.8 3.0 9.7 3.5" to 6.4" 75.7 10.1 24.2 6.5" to 9.4" 43.8 15.0 30.0 9.5" to 12.4" 25.4 16.5 29.1 12.5" to 15.4" 14.7 15.5 24.9 Total 290.4 60.0 118.0 enhancement is also important since it provides a source of income for various tribes (Woconda, pers. com.). Throughout the Region the application of a silvicultural treatment is limited by the market conditions for the resultant commercial products. Commercial fuelwood purchasers need to be able to profit from their enterprise and many factors can influence this, such as the amount of the volume removed within a given area, the local species preference, average tree diameter, slash require- ments, and access to markets. On the other side of the coin is whether on not you have quality con- tractors. The success of implementing a treatment hinges upon contractor performance. All of these are considered during project planning. Desired Diameter Distribution In today's political climate, there is a growing concern over the management of all National Forest lands. The negative appeal of even-aged management techniques has created the need for us to look at uneven-aged management techniques more seriously. The recommendations for the northern goshawk encourages group selection treatments to the ponderosa pine and mixed- conifer cover type. In the pinon-juniper type, there has been many applications of single-tree selection and group selection, however, few of these have closely followed a strict application to balance the desired diameter distribution. As a result, it is difficult to determine what is truly sustainable. In light of this, the diameter distribution repre- sented in table 1 is recommended from the Heber Study plots. This distribution is based upon a Q of 1.2 (computed on a 1 inch diameter class), a top diame- ter of 15 inches with the upper limit being flexible depending upon the available trees and a corre- sponding adjustment to the trees per acre of the top diameter class, and a residual basal area target of approximately 60 square feet. For implementa- tion, the residual trees per acre can be imple- 207 merited on a three inch diameter class, making five classes. It should be noted that this distribution is at a stand density index of 25 percent of maximum for pinon-juniper. This considered to be the onset of competition for trees. If the desired condition is to maintain a strong forage component, a lower re- sidual density would need to be selected. ADAPTIVE MANAGEMENT The highest success in applying silvicultural practices in pihon-juniper has occurred on the more productive sites. On poorer sites, the appli- cation may not be as successful. On the Carrizo Project, the initial tliinnings over a larger area are being applied at a very light intensity. The intent is to start the restoration of the interspaces. The subsequent treatments are intended to be progres- sively more intense. This approach allows manag- ers to change direction as new information and ideas are developed. In conclusion, management should be viewed as an adaptive process: we learn about the poten- tials of natural populations to sustain harvesting mainly through experience with management itself, rather than through basic research or the development of general ecological theory (Walters 1986). The challenge is to be creative, innovative, and willing to take acceptable degrees of risk in designing and testing new silvicultural practices (Bergsvik 1994). REFERENCES Balda, R.P 1987. Avian impacts on pinon-juniper woodlands. In: Everett, R.L., compiler. Proc. Pinon- juniper conference; 1986 January 13-16; Reno, NV. GTRINT-215. Ogden, UT: USDA Forest Service, In- termountain Forest and Range Experiment Station. Bergsvik, Karl. 1994. "Ecosystem Management and Silviculture". USDA-Forest Service, Unpublished Document. 4 p. Evans, R.A. 1988. Management of pinon-juniper woodlands. USDA Forest Service, Intermountain Forest and Range Experiment Station. GTR INT-249. Gottfried, Gerald and Severson, Keith. 1993. Distribu- tion and Multiresource Management of Pinon- juniper Woodlands in the Southwestern United States. In: Aldon, E.F., and D.W. Shaw, tech. eds. Proc. Managing Pinon-juniper Ecosystems for Sustainabil- ity and Social Needs. GTR RM-236. Fort Collins, CO. USDA Forest Service, Rocky Mountain Forest and Range Experiment Station. Holland, CJ. 1989. Pinyon-juniper Management in Region 3. In: Proc. of the National Silviculture Work- shop, Silvicultural Challenges and Opportunities in the 1990's, Petersburg, AK, July 10-13, 1989. USDA Forest Service, Timber Management, Washington Office. Meeuwig, R.O. 1983. Stand Dynamics and Management Alternatives for Pinyon-juniper Woodlands. In: Managing intermountain rangelands-improvements of range and wildlife. GTR INT-157. Ogden, UT. USDA Forest Service, Intermountain Forest and Range Experiment Station. Ronco, Frank. 1987. Silviculture of Woodlands and Riparian Forests of the Southwestern United States. Unpublished Report. USDA Forest Service. 1993. Watershed Management Practices for Pinon-juniper Ecosystems. Walters, Carl. 1986. Adaptive Management of Renew- able Resources. 208 Trial Applications of Low-Impact Herbicides for Pinon-juniper Control in the Southwest1 Douglas Parker2, Max Williamson3, Richard Edwards4, and Russell Ward5 Abstract. — A significant need exists to develop effective, efficient, safe, and environmentally sound approaches to control pinon-juniper trees, especially for small trees, those less than 6 feet in height. To meet this need, a new approach using a low-volume application of the herbicides Tordon 22K (picloram) and Spike 80W (tebuthiuron) in water was evaluated. Herbicides mixtures were applied to the base of selected trees just above the ground using a backpack sprayer. Tordon 22K was tested at 10, 20, 40, and 80 percent concentrations in water and Spike 80W was tested at 0.5 and 1.0 pounds of product in one gallon of water. This low-volume approach was selected to allow applicators to carry sufficient product and carrier into rough and remote areas. The goal was to only treat selected trees and avoid adversely affecting grasses and other nearby plants. The results of trial application done on the Gila and Lincoln National Forests show that satisfactory control of small trees can be achieved. The approach also may be useful for control of larger trees under selected circumstances. INTRODUCTION Pinon-juniper (P-J) woodlands occupy a vast acreage in the Southwest. Over the past 100 years, tree densities have increased many fold and trees have spread into adjacent ecosystems. This situa- tion has and will continue to cause numerous ad- verse environmental and social effects. Although considerable controversy exists over the causes of the problem and past control measures, it appears there is general agreement that on-the-ground management actions are needed to restore deterio- rating ecosystems and enhance protection of areas that will be in an unsatisfactory condition in the near future. An enormous opportunity exists to create and maintain healthy P-J ecosystems. The 1AII pesticides must be registered by appropriate State and/or Fed- eral agencies before they can be used. Pesticides can be injurious to humans, domestic animals, desirable plants and other wildlife - if they are not handled properly. Use all pesticides selectively and carefully. Follow recommended practices for the disposal of surplus pesticides and pesticide containers. 2 Entomologist, USDA Forest Service, Southwestern Region, Albu- querque, NM. Private consultant on vegetation management, Marietta, GA. 4Forester, USDA Forest Service, Lincoln National Forest, Smokey Bear Ranger District, Ruidoso, NM. 5 Range/Wildlife/Watershed Staff Officer, USDA Forest Service, Gila National Forest, Mimbres Ranger District, Mimbres, NM. key will be to have effective, safe, economical, environmentally sound, and socially acceptable methods that can be used by those responsible for managing affected lands. An excellent overview of the various manage- ment options and related recommendations is in- cluded in the publication entitled "Watershed Management Practices for Pihon-Juniper Ecosys- tems" (USDA Forest Service, 1993). Acceptable methods are available to some extent to treat larger trees, those above 6 feet in height, such as through fuelwood harvests. The control of small trees and sprouts, however, has proven to be a much more difficult problem. Hand methods (cutting, chop- ping, and grubbing) have and are being used, although the effectiveness, cost, and safety of these 209 techniques cause concern. Herbicides have been shown to be the most efficient, effective, and safe approach to control small trees, but they have not been widely used in recent years. A lawsuit in the 9 Circuit Court of Appeals and a subsequent temporary suspension of the use of herbicides on National Forest Systems lands in 1984 all but ended herbicide use on National Forests and other Federal lands in Arizona and New Mexico. This legal bar- rier no longer exists. In addition, certain herbicide formulations and application methods developed and used in the 1970's and 80's do not meet the needs of current resource managers. The high cost of compliance with the National Environmental Policy Act (NEPA) and the public controversy over the use of herbicides also contributed to the lack of herbicide use. In 1993, we decided to take another look at the herbicide option for control of unwanted trees. A review of herbicide performance showed that picloram and tebuthiuron were the most promising products, since they were broadleaf and brush products, registered for range and pasture uses (Johnsen 1987; McDaniel and White-Trifaro 1987). A positive aspect for both of these herbicides is that they will have minimal effects on established grasses at application rates specified on their re- spective labels. It should be noted that both of these herbicides are considered ''soil active" mate- rials; however, we decided to test a new application approach of applying the herbicide directly to the base of target trees. A major objective of the herbi- cide evaluation was to attempt to develop a low-volume, selective application technique that could be effectively used to treat small trees, espe- cially in rough and remote terrain, where other alternatives would not be appropriate. Broadcast applications and directed foliar sprays were ex- cluded because of environmental concerns and op- erational constraints, especially the need for a large volume of water. APPLICATION APPROACH Low-volume, basal applications of Tordon 22K (picloram) and Spike 80W (tebuthiuron) mixed with water were evaluated. Tordon 22K is a liquid formulation and Spike 80W is a wettable powder. The herbicide mixtures were applied to the base of selected trees at ground level using a backpack sprayer with a diaphragm pump (fig. 1). A model 30 gunjet with a 0002 or DE-2 spray tip was used. This equipment is relatively inexpensive; the total cost being about $125. The major advantage of the Figure 1. — Backpack sprayer used to apply herbicides to the base of trees. backpack sprayer is that the pump is sealed at the factory and tested to 70 psi to prevent leaks. The herbicide mixtures were not applied to the soil; rather, they were applied to the base of trees and sprouts just above the ground. Tordon 22K was tested at 10, 20, 40, and 80 percent concentration in water and Spike 80W was tested at 0.5 and 1.0 pounds of product in one gallon of water. These different concentrations were tried to determine the lower limit of the product that might yield sat- isfactory control. Larger trees were also treated to determine if treatment success could be achieved. A band from 2-4 inches was sprayed on these trees. The widest band of herbicide mixture was applied to the largest trees and the band was reduced pro- gressively as the diameter of the trunk decreased. Initially, the spray mixture was herbicide and water; however, a silicone wetting agent was added to the mixture beginning in March 1994 to increase the movement of the herbicide down the stem. The goal of the basal application technique was to only treat selected trees and not affect desir- able trees and shrubs within two to three feet. The low-volume approach was selected to allow appli- cators to carry sufficient product and carrier into remote areas to optimize application efficiency. 210 The tests were done on the Gila National For- est, North Star Mesa, and the Lincoln National Forest, Carrizo Demonstration Project site. Pirion pines (Pinus edulis) were treated at both sites; however, the juniper species were different. Alliga- tor juniper, Juniperus deppeana, a sprouting species, was the target on the Gila National Forest, and one-seed juniper, /. tnonosperma, was the primary species treated on the Lincoln National Forest. Permanent rectangular-shaped plots were estab- lished and marked with stakes and flags to enable evaluation of treatment effects in the future. The applications were done by walking through each plot spraying selected trees. Treatments were ap- plied in August and November, 1993, and March, 1994. Evaluations of effectiveness were made in November, 1993; March, 1994; and June, 1994. Additional evaluations are planned over the next six to nine months. TREATMENT RESULTS The results of the trial applications for both herbicides are very promising. After only three months, significant browning of foliage was observed in over half of the treated alligator juniper sprouts, seedlings and saplings of both juniper species, and pinon saplings. Larger trees showed some effects, although to an extent less than the smaller trees. Treatment effects were more apparent with the picloram mixtures, but tebuthiuron is a slower acting herbicide and the results are expected to improve over time. Treatment success continued to increase after six months, but few effects were observed on nearby untreated trees and shrubs. By nine months, foliage brown-out of trees on the Gila National Forest site exceeded 90 percent for the Tordon 20, 40, and 80 percent concentrations (Table 1). A 10 percent mixture of Tordon, which was applied six months previously, controlled about 70 percent of alligator juniper sprouts and seedlings. Since the levels of tree brown-out on the Lincoln National Forest are similar to that observed on the Gila National Forest, a separate table was not included. The mortality of pinon exceeded that of junipers on both treatment sites. On the Lincoln National Forest, exceedingly high treatment effects were observed on larger trees. For trees less than 6 feet in height, the 20 percent Tordon mixtures produced control results as good as the 40 and 80 percent concentrations. More time is needed for to evaluate the 10 percent mixture, but it appears the lower limit that will yield satisfactory results could be below the 20 percent concentration level. Thus, the current cost for the herbicide will be somewhere in the range of $8-16 per acre. Also, the silicone wetting agent appeared to improve the movement of the spray mixture down the stem, which could be a critical factor when lower concentrations of herbicide might be used. Although the Spike 80W treatments have not produced tree brown-out as high as that observed with Tordon 22K, the overall treatment success may be similar in a few more months. As shown in Table 1, the half pound per gallon of water mixture of Spike 80W yielded better initial foliage brown-out than the pound per gallon mixture. This was probably due to the higher volume of the spray mixture applied to each tree. The results on the Lincoln National Forest are similar, except no sprouting stumps were treated. It will take more time before the final conclusions can be made about the effects of the various treatments, especially for the larger junipers. Table 1.— Pihon-juniper herbicide trail results, Gila National Forest, North Star Mesa, Evaluated on June 1, 1994. Foliage Brown-out (percent) Plot No. Date Treated Herbicide Sprouts Seedlings/Saplings Larger Trees 1 8/23/93 Tordon 22K20% 90+ 90+ 20 2 8/23/93 Tordon 22K 40% 90+ 90+ 20 3 8/23/93 Tordon 22 K 80% 90+ 90+ 30 4 8/23/93 Spike 80W (1/2 Ib./gal.) 60 60 10 5 8/23/93 Spike 80W (1 Ib./gal.) 40 40 5 6 11/15/93 Tordon 22K10% 70 70 10 7 11/15/93 TORDON 22 K 20% 80 80 10 8 3/7/94 Tordon 22K20% 80 80 10 9 3/7/94 Spike 80W 1/2 Ib./gal.) 30 30 5 211 DISCUSSION It appears that the spray mixtures flow down the stem following root profiles and the herbicide is absorbed by root hairs around the base of treated trees. Susceptible trees and shrubs within two to three feet of treated trees have shown little signs of herbicidal activity. Besides being selective, a major benefit of the basal application approach is that the herbicides are not applied to the soil. The next step will be to focus attention on how to utilize this low-volume, selective application approach to achieve desired future conditions. A few possible circumstances where this new tech- nique may be useful follow: Create or Maintain Existing Openings. Treat- ment of undesirable or excess seedlings and saplings to create or maintain existing openings in P-J stands would be one of the most eco- nomical uses of this new application technique. It probably would cost less than mechanical grubbing, and a major advantage would be to avoid ground disturbing activities within pro- tected areas. Fuelwood Harvest Areas. Treatment of alligator juniper sprouts and excess seedlings and sap- lings would be another economical use follow- ing fuelwood harvest. This approach would be particularly desirable in areas where prescribed fire might not be an option. Thinning. Thinning in P-J woodlands has been done to promote growth of remaining trees, release understory ground vegetation, and most recently, to create fuels to enable the use of prescribed fire. The low-volume, selective herbicide technique would be appropriate where trees are too small to be of commercial value or in rough and remote terrain. Removal of small trees that occur under larger trees would help to prevent fire from moving into the crown of the larger trees that are consid- ered to be desirable. Wildlife Openings. Large expanses of dense canopied P-J woodlands are common in the Southwest and offer little in terms of plant and animal diversity. A directed basal herbicide treatment could be used to create openings in selected areas across the landscape to increase diversity, release native plants, and reduce sheet erosion. Dense stands of smaller trees in inaccessible areas would be most suited for this approach. Creation of Snags. Snags are important habitat for several birds and other wildlife. Creation of snags may be one of the few economical uses of herbicides to treat large trees. Areas that are in- accessible would be best suited because snags are usually harvested by the public in accessible areas. Protection of Archaeological and Historic Sites. In some cases, these sites can be ad- versely affected by erosion and arroyo cutting and may need to be protected through water- shed improvement efforts. Since mechanical treatments can damage these fragile resources, management options are often limited. Herbi- cides may offer an attractive option to restore P-J woodlands to a healthy condition to provide prolonged protection of these sites. CONCLUSIONS Ecosystem management is a concept that will guide vegetation management on public lands in the future. Ecosystem management involves using an ecological approach to achieve multiple-use ob- jectives by blending the needs of people, environ- mental values, and scientifically based techniques. It must be realized that P-J ecosystems, which are threatened or are being damaged by an "over abundance of trees", can only be restored to a healthier state through the removal of trees. All available methods — fuelwood harvest, mechanical methods, prescribed fire and cultural practices, like limiting livestock grazing — will need to be used. Herbicides will be one of these tools to achieve management objectives, used individually or in combination with other methods, as part of an in- tegrated vegetation management approach. There has been a wealth of information to show that herbicide formulations used in modern re- source management are "safe" when used prop- erly. Herbicides are among the most rigorously tested consumer products on the market today. Before they are registered for use, herbicides must meet strict standards of human health protection and environmental safety. In addition, a major benefit of the herbicides evaluated in this study is that they provide selectively through both directed application and the inherent selective nature of the products. It will be particularly important to have 212 thorough environmental analyses for projects on Federal lands, which incorporate available risk assessments, to respond to possible challenges to proposed herbicide projects. It is important to note that risk assessments developed in recent years, such as the Risk Assessment For Herbicide Use in the Forest Service Regions 1,2,3,4, and 10 and on Bonneville Power Administration Sites, which includes the Southwest, have withstood legal challenges. As always, well trained applicators and special- ists will be needed to ensure program success. Significantly, the public will demand that resource managers and applicators be knowledgeable about the methods and herbicides that are proposed for use. Comprehensive training and certification programs will need to be developed and imple- mented before this new herbicide approach can be used to help solve the enormous environmental and social problems that are occurring in P-J eco- systems in the Southwest. LITERATURE CITED Johnson, T.N. Jr. 1986. Using herbicides for pin- yon-juniper control in the Southwest. E330-334. In: Everett, R.L., compiler. Proc. Pinon-Juniper Confer- ence. January 13-16, 1986, Reno, NV. GTR INT-215. Ogden, UT: USDA Forest Service. Intermountain Re- search Station. McDaniel, K.C. and L. White-Trifaro. 1986. Selective control of pinyon-juniper with herbicides, p. 448-455. In: Everett, R.L., compiler. Proc. Pinon-Juniper Con- ference. January 13-16, 1986, Reno, NV. GTR INT-215. Ogden, UT: USDA Forest Service. Intermountain Re- search Station. USDA Forest Service. 1993. Watershed management Practices for Pinon-Juniper Ecosystems. Southwest- ern Region, Albuquerque, NM. 213 Pinon-Juniper Fuelwood Markets in the Southwest Lawrence A. Schmidt1 Abstract. — This study estimates the commercial harvest and sale of pinon- juniper (pj) fuelwood from state, private, and public forests in Arizona, Colorado, New Mexico, Nevada, and Utah. The demand for fuelwood peaked five or six years after the 1973 oil embargo. Reports and personal interviews suggest there is less fuelwood consumed today than there was during early 1970 and 1980. There is a strong public awareness for better air quality. Many cities and towns have "no-burn" days during winter temperature inversions. Persons residing in these communities cannot use the same amount of fuelwood they used in the past, before the restrictions were in effect. A surge of interest by the sagebrush rebellion groups has seemingly reduced availability of fuelwood from public and some state woodlands. INTRODUCTION Information contained in this paper came from Fiscal Year-92 and 93 annual reports prepared by federal, and state forest officers. Woodland and timber acreage are from the Rocky Mountain Inter- Mountain Research Station publications listed in the bibliography. During the last ten years fuelwood harvests have varied from year to year (Table 1). One location may increase its harvest while another experiences a decrease. Looking at values for the Southwest, fuelwood use is less than it was twenty years ago. People throughout the United States are concerned about air quality, and the Southwest is no exception. During winter months there are several days and sometime weeks of weather that produce temperature inversions. Communities with clean air ordinances prohibit the use of wood stoves and fireplaces without catalytic converters when there are temperature inversions. Typically radio and TV stations announce no burn days to keep the public informed. The Colorado front range has strict regulations concerning the use of conventional wood stoves Table 1 .—Estimated Timber and Woodlands (Millions of Acres) Southwestern U.S. State Timber Woodlands Arizona 72.8 5.5 9.1 Colorado 66.6 15.0 6.0 New Mexico 77.0 6.2 9.0 Nevada 70.0 0.7 9.0 Utah 52.5 16.0 9.0 Total 338.9 43.4 42.1 and fireplaces. A few cities will not allow contractors to build homes with fireplaces. When inversions occur residents burning fuelwood must switch to cleaner burning fuels such as natural gas or LP fuels. Approved wood burning appliances may include wood pellet stoves or stoves and fireplaces with catalytic converters. Since the oil embargo, natural gas and liquid petroleum gas prices have fluctuated, although the cost of these fuels in normally less than fuelwood. Today most homeowners know that petroleum fuels are less expensive, cleaner burning and more readily available than fuelwood. People that have wood stoves and natural gas furnaces prefer wood heat. Wood has many desirable characteristics such as aroma, flame lengths, color and sound. Forester, Bureau of Indian Affairs, Phoenix Area Office. 214 TRIBAL WOODLAND ACREAGE The Bureau of Indian Affairs (BIA) is the trustee for most Tribes in the Albuquerque, Navajo, and Phoenix Areas that have significant PJ acreage. The Albuquerque Area is responsible for tribes in New Mexico, and Southwestern Colorado namely Southern Ute, and Ute Mountain. Albuquerque Area has 1,400,926 woodland acres, and 879,641 acres are commercially accessible. Navajo tribal forestry has a woodland forester that works on the Navajo Reservation and the new lands. New lands are part of the Navajo Reservation created by congress, to relocate Navajo people that previously lived on the Navajo and Hopi joint use area. The Navajo tribe oversees almost four million acres of woodlands, the largest woodland acreage of any tribe in the nation (Table 2). There are fourteen reservations in the Phoenix Area that have 2,089,162 acres of woodland with 752,502 acres classified as commercially accessible. This does not include the arid, lower elevation desert reservations with significant stands of mesquite. This includes Tohono O'odham, Gila River, Fort McDowell, and Tribes along the lower Colorado River. Range Conservationists at Tohono O'odham estimate they have almost one million acres of mesquite. Most of this mesquite is too small for fuel wood. Many tribes have annual resolutions and ordinances for fuelwood and small forest products. These documents set prices for miscellaneous forest products, and develop guidelines for harvest. Tribal members have cut fuelwood for many years to supplement their income. Tribal members must pay for a paid permit, which allows them to sell on or off their reservation. Alligator juniper (Juniperus deppeana) is a preferred species. When dry it is lighter to work with, has a pleasant aroma, and except forks and large knots, it is easy to split. Indian cutters try to sell all of their wood to tribal members or merchants and people in towns close by. Fuelwood prices range between $50 and $125. Factors that cause variability includes: the Table 2. — Woodland Acres (millions) and Estimated Allowable Cut(mmbf) Includes tribal ownerships in Arizona, New Mexico, Colorado, Nevada, and Utah Table 3.— Average wholesale fuelwood prices delivered to broker. State Gambel Oak Pihon Mesquite Alligator Juniper Other Juniper AM \L\J\ la •poo $85 New Mexico $80 $105 $80 $70 Colorado $65 $55 $55 $50 Nevada $90 $60 $55 Utah $85 $50 $45 So. Calif. $170 $100 $175 $110 $95 Area Acres Allowable cut Albuquerque 1.4 2.056 Navajo 4.0 1.500 Phoenix ZA. 4.033 Total 7.5 7.589 season of the year, availability, road conditions, competition and the purchaser. Preferred clients include motels managers that are willing to pay retail for one or more cords. Brokers buy in large quantities they pay the least amount, since they have their own cutters or have contracts with other operators (Table 3). Larger brokers typically purchase in the spring and summer to have a large inventory before winter. This practice allows them enough time to split and dry the wood before selling to their distributor. Overall there has not been too much of a change in prices operators receive for fuelwood during the last four years. Operators that sort large trees can create value added items to increase their profit. Junipers with 6/-8' straight stems make excellent fence and corner posts. The return for posts is much greater than cutting these same straight pieces into fuelwood. Honey mesquite (Prosopis glandulosa) is worth more than a dollar a pound. Mesquite is a favorite for Southwestern furniture makers. To produce lumber for furniture stock, mesquite trees need to be at least 10 inches in diameter, and produce boards four feet or longer. In addition mesquite should be clear or have small defects. Furniture makers will pay $5.00 a board foot for quality mesquite. TECHNIQUES THAT HELP SAVE TIME AND PRODUCE MORE FUELWOOD Bennett estimates that non-commercial cutters spend the equivalent of two hundred dollars for a cord of fuelwood. A novice cutter would be wise too pay attention to this estimate. Hard working cutters with several years of experience can reduce that estimate to one hundred dollars or less. Not all fuelwood cutters produce the same amount of wood each day. There are many vari- ables. Knowing where to find a sustainable supply of fuelwood to selectively harvested without harming the ecosystem is necessary. Sawyers need to have systems developed that will maximize their productivity. Having extra sharpened chains, and 215 additional chain saws in case one breaks down will make a difference in having a profitable day or going home disappointed. The largest fuelwood operators cut thousands of cords a year to meet their contractual obligations. Arizona has five or six large commercial fuelwood operators (3,000+ cords) that package cubic foot bundles using plastic bags or stretch wrap. These are the packages people see every day displayed at convenience and many other stores throughout the state. The largest operators normally hire their own crews and pay them on a piece count basis. One large operator has a contract to cut pihon pine and one seed juniper for a fee of $9.00 a cord. By the time a cord of fuelwood is hauled from the woodlands job-site into Phoenix, most operators have $90-100 invested. The crew doing the cutting and loading receive approximately $50.00 per cord. Hauling fuelwood from the Flagstaff-Seligman area to Phoenix costs between $20-$25 a cord. Transportation variables include the size and condition of the truck, distance hauled, road conditions, and whether the fuelwood is dry or green (Table 4). Several truck brokers will not haul fuelwood because of bad experiences encountered in the past. Distributors prefer to haul wood bundles packaged and loaded on pallets. One buyer in California will buy loose blocks, if it is loaded on flat bed trailers with four foot side racks. An operator at Lakeside, AZ has a method of loading loose blocks of wood into plastic bags that serve as pallets. This enables him to use a forklift to lift the bags into truck trailers. One disadvantage is the cost of the bags, they vary between $17 and $27 compared to $3-$5 for wooden pallets. An important advantage is that after the loose pieces are loaded by hand they can be loaded with a forklift. Loading by a forklift saves many hours of hand labor. These plastic bags are made out of a strong poly material and can be used several times. Some of the largest operators that depend on ranchers for their wood supply are concerned about future fuelwood sources. One broker estimates he has three or four years supply remaining (Table 5). Table 4. — Transportation costs truck and rail. Truck Rail San Carlos- Phx. $275 — San Carlos- L. A. $550 $1 .24/1 00# load at Phx. Hondah- Phoenix $475 Hondah-LA. $750 $1 .24/1 00# load at Phx. Table 5. — 1993 Fuelwood harvest thousand board feet (MBF). Land AZ. CO. NM NV UT Total BIA 7,307 1,994 8,923 256 18,480 BLM*(1992) 587 4,174 3,047 4,381 12,189 PRIVATE 2,200 4,100 3,400 200 500 10,400 STATE 300 675 234 1,209 USFS 25.400 21.300 16.800 1,358 14.455 79.313 TOTAL 35,794 32,243 29.123 4,505 19,826 121,591 DIFFERENT METHODS OF PACKAGING FUELWOOD The largest fuelwood operator in Arizona processes ponderosa pine (Pinus Ponderosa) from Forest Service multi-product sales. He also harvests PJ from a ranch in Northern Arizona. His Arizona operation including PP and PJ processes approximately 7,000 cords a year. His crews also cut 5,000 cords of oak and other species from private land in California. This operation packages fuelwood bundles by stretch wrapping rather than using plastic bags. The fastest wrappers can wrap one thousand cubic foot bundles in an eight-hour shift. They place 40 packages per pallet or three pallets per cord. A van will hold 14 to 18 cords or 1680 to 2160 cu. ft. An average load consists of two thousand bundles loaded on 50 pallets. When packaging fuelwood it is important to make sure there is enough wood in the package. Packaged fuelwood normally has labels indicating the kind and amount of wood they are buying. Weights and measures inspectors check packages to verify that the consumers are getting what they pay for. Many processors will intentionally load a bit more than required to pass weights and measures inspections. Another technique is to sell .75 cu. ft. packages. Smaller packages create more work, but the average consumer generally does not notice any difference. All fuelwood operators need a means of distributing their product. A large grocery distribution company near Phoenix delivers most of the packaged fuelwood to convenience and other stores throughout Arizona plus Las Vegas, Nevada. This company pays about $2.00 per bundle for packaged fuelwood they deliver. This distributor will not accept stretch wrapped packages of fuelwood, because stretch wrapped bundles are open ended. This type of bundle can result in small amounts of bark and dust spilling on floors. 216 Plastic bags cost about 20 cents each depending on whether the bag is clear plastic or has a printed logo. A few years ago shrink wrap was the favored packaging medium. This system uses heavy plastic for packaging. Next these packages are placed on a conveyor belt and fed into an oven. Heat from the oven causes the heavy plastic to shrink tightly around the fuelwood. Shrink wrapping is slower and more expensive than stretch wrap or plastic bags. Other packaging operations prefer cardboard boxes, because they stack better than plastic and contain twice as much wood. Cardboard boxes are much more expensive than plastic bags or wrap. All operators mentioned uniformity, quality and reliability as the most important items to be successful in the fuelwood business. An established dealer in Northern New Mexico believes the key to any fuelwood business is consistency. Fuelwood cut to uniform lengths, limbs trimmed, and sold in quantities as agreed. A fuelwood operator will not stay in business if they attempt to sell green wood during the winter burning season. The product must be consistent day in and day out. Because it is easy to start a fuelwood business the turnover rate of operators is very high. It is unusual to find operators that have been in business more than five or six years. This business is extremely competitive. Intense competition keeps wholesale prices low. Anyone can sell a few cords at retail prices, however; this will not be the case when selling hundreds of cords. The fuelwood business may look attractive, in reality most that venture the business go broke. Many hard working operators have given up because they are unable to locate reliable supplies of material. Although the amount of fuelwood harvested has dropped, the demand exceeds the supply. When there is a limited supply of fuelwood federal and state resource manager's generally give priority to individuals rather than commercial operators. SORT LARGE WOOD TO INCREASE PROFIT When harvesting fuelwood it is easy to concentrate on a single output. Depending of the size, quality, and straightness of trees it is wise look at making other end products besides fuelwood. By sorting material more cash is generated by selling posts, poles and lumber. For example there is a demand for mesquite large enough to make lumber for furniture, bowls, and cutting boards. Mesquite boards 12" x 8" by 3/8" thick sanded and finished retail for $12. There are many value added items made out of wood. Craftsmen turning mesquite is a big business in Texas. The good thing about most curio businesses is that they affordable. By starting small, craftspeople build their business to a size that meets their needs. Although PJ harvesting has decreased there is still a large amount of material being used for fuelwood and other products. There will always be significant quantities of PJ available. Today there is less PJ available from public lands. This may not be the case a few years from now. There is a good prospect that several new businesses will emerge in the future that will use pifion-juniper. LITERATURE CITED Bennett, Duane A. 1990 Fuelwood Extraction in Southeast Arizona. In: 1992 Proceedings of the Oak and Associated Woodlands symposium, April 27-30 1992, Sierra Vista, Arizona. Benson, Robert E. Green, Alan W. 1987 Colorado's Timber Resources. Res. Bull. INT-48. Ogden, UT; U.S. Department of Agriculture, Forest Service, Intermountain Research Station: 53 p. Born, J. David; Tymcio, Ronald R; Casey, Osborne E. 1992. Nevada forest resources. Res. Bull. INT-78. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station. 64 p. Bureau of Indian Affairs. 1988. Native American woodland resources: A national overview, Assessing the resource potential land management needs. U.S. Department of Interior, Bureau of Indian Affairs, Branch of Forest Resources Planning. Portland, OR: 139 p. Connor, Roger C; Born, J. David; Green, Alan W. 1988 Colorado's Woodland Resources on State and Private Land. Res. Bull. INT-50. Ogden, UT; U.S. Department of Agriculture, Forest Service, Intermountain Research Station: 50 p. Conner, Roger C; Born, J. David; Green, Alan W; O'Brien, Renee A. 1990 Forest resources of Arizona. Res. Bull. INT-69. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station: 92 p. Larson, Robert E. 1988 An Investigation into Fuelwood Wholesale/Retail Outlets. Report prepared for Hualapai Tribe regarding a fuelwood enterprise: 11 p. Larson, Robert E.; Bhoopal, Chan. 1988 Arizona Forest Products Directory 1987-1988 Northern Arizona University, College of Forestry; Flagstaff, AZ: 57 p. McLain, William H. 1988 Arizona's 1984 Fuelwood Harvest. Res. Bull. INT-57. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station: 8 p. 217 Miller, Ronald K.; Albert, Steven K. 1993. Zuni cultural relationships to pinon-juniper woodlands. In: Aldon, Earl F. and Douglas W. Shaw, technical coordinators. Proceedings—Managing pinon-juniper ecosystems for sustainability and social needs; 1993 April 26-30; Santa Fe, NM. Gen. Tech. Rep. RM-236. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station: 74-78. Van Hooser, Dwane D.; Green, Alan W. 1983 Utah's Forest Resources, 1978. Res. Bull. INT-30. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station. 58 p. Van Hooser, Dwane D.; O'Brien, Renee A.; Collins, Dennis C. 1993 New Mexico's Forest Resources, Res. Bull. INT-79. Ogden UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station. 110 p. 218 Ecosystem Management Research in an "Old Growth" Pinon- Juniper Woodland William H. Kruse and Hazel M. Perry1 Abstract. — The Rocky Mountain Forest and Range Experiment Station of the USDA Forest Service is conducting research in pinon-juniper wood- lands in cooperation with the Heber Ranger District, Apache-Sitgreaves Na- tional Forest. This paper describes the study, objectives, preliminary results since the study began in 1989, and anticipated results for when the study is concludes in 1999. INTRODUCTION Some pinon-juniper woodlands near Mud Tank on the Heber District of the Apache-Sitgreaves National Forest are being harvested for fuelwood. These woodlands, which we tenuously call "old growth" woodlands, appear to be older than 200 years, appear to have reached maximum growth, and appear to be mature and stable. Because J> 90% of the overstory was cut by 1993, this "old growth" condition has been extremely modified. Previous research (Arnold et al. 1964; Clary and Jameson 1981) has shown that the understory biomass often increases with overstory modifica- tion, but often quality forages and/or other impor- tant plant species do not respond as expected. Hence, in these situations, site quality is compro- mised if the anticipated quality understory biomass response is not increased or improved from over- story removal. What specifically are the effects of overstory removal on slash, small mammal populations, and the nutrient bases? Are important nutrients being removed from the site with the fuelwood? How do the different plant communities and species populations respond — especially elk, deer, and livestock forage — to various treatments and to the nutrient regimes? Slash, scattered and left on the site, is an impor- tant nutrient storage medium for slow release through decomposition into the site. Should the long-term productivity of the site be sustained as overstory production (for fuelwood and other 1USDA Forest Service, Rocky Mountain Forest and Range Experi- ment Station, Flagstaff, AZ. wood products) or understory (as forages for live- stock and wildlife)? Or are both possible by man- aging the understory for the short term and over- story products for the long term? How do the nutrient cycles affect these management options, and how does burning versus leaving the slash affect the nutrient cycles in both the short and long term? Stable ecosystems are generally sustainable and, following disturbance, generally revert to the most stable condition. In the Mud Tanks Study, ecosystem dynamics are being followed in three areas: 1) understory/overstory relationships related to habitat type, 2) influence of nutrients and nutri- ent cycling relative to site productivity, and 3) relationship of the habitat type to the fauna, specifically small mammals. Small mammals are being studied because of their importance in the food chain and as an indicator of site productivity. Forage bases are also being evaluated for the larger ungulates. Evaluating these three areas will pro- vide a basis to move from monitoring successional stages to managing the ecosystem. The overstory/understory relationships in- clude three overstory conditions; 1) complete removal, 2) commercial fuelwood removal only, leaving an advanced form of regeneration, and 3) a silviculture treatment, single-tree selection. Overstory species' regeneration following treat- ment will be studied simultaneously in the Mud Tank Study as it relates to available nutrients and slash and its affect on small mammals and forage production. 219 OVERSTORY/UNDERSTORY RELATIONSHIP STUDIES Austin (1987) presented data showing little suc- cessional change in a plant community within a mature pinon-juniper woodland during a ten-year period, suggesting a climax ecosystem. Following disturbance, his research shows a constant reduc- tion of understory vegetation on sites as soon as tree species become reestablished. At least some pinon-juniper ecosystem structures appear quite responsive to recovery mechanisms, further sug- gesting stability. When a forest or woodland ecosystem is in equilibrium, the primary flow of energy and nutri- ents is to maintain and sustain the overstory com- ponent (Smith 1966). The ecosystem appears most stable, displaying the least change, when the over- story is "old growth" or "most" mature. When changes or modifications are made to the overstory, the understory responds significantly in terms of increased understory biomass production and plant diversity (Arnold et al. 1964). Site productiv- ity is further affected by the disposition of the slash (lop and scatter, burning, crushing), and the re- moval of the fuelwood nutrient base. What then, is the relationship between the overstory and the understory as it relates to site productivity? How is site productivity described? An important aspect of the Mud Tanks research is to examine the understory plant community following disturbance to the overstory. Past re- search (Arnold et al. 1964) shows a definite increase in the quantity of the understory biomass produc- tion, but past research results (Clary et al. 1974) are mixed with regard to a change in the quality of the site which, in the past was a reflection of a sustain- able, marketable product such as increased live- stock forage production for beef or water runoff into reservoirs. Study Area and Background The Mud Tanks Study was initiated in 1989 on 33, 4.0 ha study units. Understory production data were taken on all study units. Small mammal live trapping began in 1990 on 16 units. The taking of Table 1. — Total soil nutrients from Mud Tanks Study Area. % ppm Nitrogen 1.82 Magnesium 0.10 Zinc 18.82 Phosphorus 0.71 Potassium 1.09 Manganese 50.82 Calcium 0.42 iron 0.02 overstory inventory, pre-harvest slash inventory, and soil/plant nutrient data commenced in 1990 and was completed in 1991 on 30 study units. These preliminary data were all pre-harvest inven- tory data and represent the untreated or natural condition. The first harvesting began in December 1991, and the data from 1992 and 1993 reflect some harvesting effects. Nutrient Cycling Studies The nutrient cycling component of this re- search examines soil-plant nutrient relationships and the cycling of the nutrients, following a series of overstory treatments, principally fuelwood harvest. Specific objectives are to determine changes in total and available nutrients associated with different overstory regimes assessing the interrelationships between nutrient concentrations in soils and plants, including slash. The research will determine effects of wood harvesting and slash burning on levels of selected nutrients in one understory species blue grama (Bouteloua gracilis) and the primary overstory species, Pinus edulis and Juniperus monosperma. Treatment effects emphasiz- ing sustainability are central to the research objec- tives in determining whether or not sufficient nutrients are available in the concentrations to adequately sustain the ecosystem for the desired future condition. Emphasis is being placed on nitrogen, phos- phorus, and carbon because of their importance in southwestern ecosystems (Clary and Jameson, 1981 Evans 1988), but several other elements (calcium, potassium, magnesium, sodium, manganese, iron, and zinc) are also included. Table 1 shows some of the preliminary results prior to harvesting. All of the measurements were made on Kjeldahl diges- tions. Nitrogen was measured on an Antek nitro- gen analyzer. Phosphorus was measured by ascor- bic acid method. The magnesium, potassium, iron, zinc, and manganese were measured by atomic absorption spectrophotometry. Total nutrient losses will be determined by comparing nitrogen, phosphorus, calcium, magne- sium, potassium, iron, zinc, and managanese on slash, blue grama, and litter prior to and following fuelwood harvest and burning. The same nutri- ents, along with organic carbon, will be measured on soil collected before and after fuelwood harvest and fire. Plant available nutrients will be assessed by measuring ammonia, nitrate, and phosphorus in the 0-2 and 2-10 cm soil depths in all plots before 220 burning, immediately following burning, and several times following fire when forage quality is sampled. Plant Nutrient Measurements Nutrient studies of understory vegetation are limited to one species, blue grama, because it is the only ubiquitous species within the study area. Plant and associated soil samples were collected from 1990 to the present and are in the laboratory for processing. By 1995, all the burning will be completed and the sampling will continue through the entire period on uncut, cut without burning, and cut with burning treatments. Selected soil and plant nutrients are being ex- amined to determine if patterns or other sequential relationships exist between concentrations of soil nutrients and those in the plants. Particular em- phasis will be given to nutrient loss via burning and biomass removal and how this relates to long- term uptake by overstory and understory plants. Again, emphasis is on those relationships that provide information on sustaining or pooling nutrients required for specific habitat type(s). Understory/Overstory Baseline Data Results Trees were measured by standard procedures on eight randomly located 0.04 ha circular plots within each 4 ha treatment area or unit (Table 2). The circular plots were measured prior to and immediately after harvest to determine levels of harvesting and non-commercial overstory survival. Slash was measured using the intercept method of Brown (1974). The null hypothesis is that different levels of harvesting, with and without slash dis- posal and burning, will not affect the survival of advance regeneration and growth, the understory plant components, and the small mammal popula- tions. A significant (p < 0.001) difference existed between years (1990 and 1991) on control and pre- treatment study units (Fig. 1). This difference existed for both blue grama (p < 0.006) and perennial forbs (p < 0.001). No significant differences (p > 0.05) were determined for the other plant classes. This increased production was probably because of an exceptionally wet spring in 1991. Also, between 1990 and 1991, no significant differences- were determined between the control Table 2. — Understory, overstory, and slash baseline data. 72 Juniperus deppeana alligator juniper J. monosperma one-seed juniper Pinus edulis pihon P. ponderosa ponderosa pine Total Basal Area m 1.8 BA 7.8 BA 3.7 BA 1.2 BA I4.5BA % Pom. 13% 54% 25% 8% 100% Slash: 8.66 tons/acre (pre-cutting) 55.71 mVha.(post) UNCUT UNITS CUT UNITS (understory production grams/meter2) 1990 1991 1992 1993 1992 1993 BG* 2.02 9.81 2.58 2.77 5.92 7.55 PG* 0.10 1.26 0.35 0.59 0.41 124 AG* 0.01 0.00 0.01 0.00 0.01 0.05 PF* 0.55 3.37 3.66 3.95 7.67 9.96 AF* 0.24 0.24 0.21 0.59 0.59 0.50 Total 2.92 14.68 6.81 7.90 14.60 19.30 BG = Bouteloua gracilis (blue grama), PG = Perennial Grass AG = Annual Grass, PF = Perennial Forbs, AF = Annual Forb and the "to be cut" units suggesting similarity between all the study units. ANOVA tests performed on fhe 1991 vs. 1992 data also showed no significant differences between the study units or years. A few study units were harvested during 1992, but effects of the overstory removal are not evident yet By 1993, however, some significant changes were apparent (Fig. 1). Highly significant (p < 0.001) differences existed between harvested and control units. These differences were found for total production, blue grama, and perennial forbs. Other plant classes — annual grass, annual forbs, and perennial grasses — exhibited little or no change. Blue grama and the perennial forbs con- tribute most toward the total production (Fig. 2). The understory production continues to be measured in every subplot by using an ocular estimate by plot technique employing double sampling (Cook and Stubbendieck 1986). Data collection commenced in September 1990 and will be continued each year thereafter, including the treatment years and three or four years post treat- ment. The data are being analyzed to test the null hypothesis that understory production does not differ among treatments. Years will be introduced into the analysis as a repeated measurement. Even- tually, relationships between overstory parameters and understory production will also be examined to determine if functional models can be developed to predict understory yield from the overstory condition of pinon-juniper. 221 TOTAL PRODUCTION Treatment (units) vs Control (units) 1990 1992 I TREAT H CONTRL TREAT CONTRL 1990 3.2 2.9 1991 15.9 14.7 1992 14.6 6.8 1993 19.3 7.9 PERENNIAL FORB PROD. Treatment (units) vs Control (units) 1990 1991 1992 1993 I TREAT I CONTRL TREAT CONTRL 1990 0.6 1.1 1991 5.4 8.7 1992 7.7 3.7 1993 9.9 3.9 BLUE GRAMA PROD. Treatment (units) vs Control (units) 10 ■s E & E CO i_ 1990 1991 1992 1993 I TREAT I CONTRL TREAT CONTRL 1990 2.4 2.1 1991 9.7 9.8 1992 5.9 2.6 1993 7.6 2.8 ANNUAL FORB PROD. Treatment (units) vs Control (units) 0.6 - 0.5 1990 1991 1992 1993 I TREAT I CONTRL TREAT CONTRL 1990 0.2 0.2 1991 0.6 0.2 1992 0.3 0.2 1993 0.5 0.6 Figure 1.— Above ground understory biomass production (g/m2) for total, blue grama, perennial, and annual forb production. Signifi- cant differences were between 1990 and 1991 (these plant classes) for production. No differences were determined between the control and the "to be harvested" units. Significant differences due to treatment were determined. Small Mammal Measurements As with the understory data collection, the relative abundance and species composition of small mammals are being evaluated. However, small mammal live-trapping is being limited to only two overstory treatments, type conversion and controls. Table 3 shows some preliminary results from the first four years of trapping (two pre-treatment years and two years that include some cut units). Trapping is being conducted dur- ing a four week period (July-August) each year. Species richness and eveness are being calculated for each treatment combination according to meth- ods described by Ludwig and Reynolds (1988). The null hypotheses of no differences in (1) total num- ber of small mammals or (2) total number of species among treatments are being tested. Average captures for the deer mouse (Peromyscus maniculatus) , pinon mouse (P. truei), all others, and total small mammal captures are represented in Table 3. These data express the dominance of the two Peromyscus species relative to the total numbers of all captured animals. A fluctuation in species diversity accompanies the annual fluctuation, however. Notable is the low number of "all others" suggesting a lack of spe- cies diversity for the area. A more detailed ac- count of these results can be found in Kruse (1994), this symposium. 222 Table 3. — Mean captures for uncut and cut treatments by year. 1990 1991 1992 1993 UC C UC C UC C UC C P. maniculatus 3.8 4.1 18.4 16.6 2.9 7.5 11.8 45.6 P. truei 16.3 12.8 20.1 19.4 12.5 10.4 30.5 20.0 All Others 30 35 6J. 55 2J. 3Z 5J> 10.7 Total animals 23.1 20.4 44.6 41 .5 17.5 21.6 47.9 76.3 UC = Uncut, C = Cut The small mammal populations fluctuated dramatically during the first four years of the study (Fig. 3). The first analysis showed significant differ- ences between years, population densities, and unit study areas (Fig. 3). A second analysis on all small mammal cap- tures showed similarity among all plots prior to harvest. This analysis showed that all "to-be" cut (not now cut) units and all control (never- to -be cut) units were similar. The "Small Mammal Captures" graph also expresses significant doubling of total population numbers from 1990 to 1991. SUMMARY Significance of the Research Results from this research will have direct ap- plication on over 12 million ha of pifion-juniper woodlands in Arizona and New Mexico and impli- cations on about 19 million ha throughout the western United States. Information gained from this study will provide a basis for developing im- proved guidelines for ecosystem management of SPECIES PERCENTAGES OF TOTALS BY YEAR ■ 1991 TRTB1991 CNT ~ 1 993 TRT B 1 993 CNT Figure 2.— Composition of understory biomass. F = annual forbs, AG = annual grass, BG = blue grama, PF = perennial forbs, PG = perennial grass. TOTAL CAPTURES Small Mammals 1990-1993 ■ ill 1990 1991 1992 1993 Year: 1990 1991 1992 1993 Mean: 43.5 86.4 39.1 124.9 +/- 6 +/- 8 +/- 8 +/- 1 8 Figure 3. — Difference between years 1990 vs 1991 (p< 0.001). Difference between study units prior to harvest, N.S. Highly significant difference between control and treatment, 1992 vs 1993 (p<0.001) (Kruse 1994). the pinon-juniper woodland in the southwestern United States. The nutrient portion of this research project is especially important in that it magnifies an unprecedented endeavor to simultaneously study the effect of fuelwood harvesting and slash management on livestock and wildlife forages, small mammal populations, understory/overstory plant relationships, and nutrient cycling. LITERATURE CITED Arnold, Joseph E; Jameson, Donald A.; Reid, Elbert H. 1964. The Pinon-juniper Type of Arizona: Effects of Grazing, Fire and Tree Control. Prod. Res. Rep. 84. Washington, DC: U.S. Department of Agriculture, Forest Service. 28. Austin, Dennis D. 1987. Plant Community Changes Within a Mature Pinon-juniper Woodland. Plant Ecology. 47 (1): 96-99. Brown, J. K. 1974. Handbook for Inventorying Downed Woody Fuels. Gen. Tech. Rep. INT-16. Ogden, UT: U.S. Department of Agriculture, Forest Service, In- termountain Forest and Range Experiment Station. 24. Clary, Warren P; Jameson, Donald A. 1981. Herbage Production Following Tree and Shrub Removal in the Pifion-juniper Type of Arizona. Journal of Range Management 34(2):109-113. Clary Warren P, Malchus B. Baker, Jr., Paul F. O'Connell, Thomas N. Johnsen, Jr., and Ralph E. Cambell. 1974. Effects of pinon-juniper removal on natural resource products and uses in Arizona. USDA For. Serv. Res. 223 Pap. RM-128, 28p. Rocky Mt. For. and Range Exp. Stn., Fort Collins, Colo. 80526. Cook, C. Wayne; Stubbendieck, James. 1986. Range Research: Basic Problems and Techniques. Denver, CO: Society for Range Management. 317. Evans, Raymond A. 1988. Management of pinon-juniper woodlands. Gen. Tech. Rep. INT-249. Ogden, UT: U.S. Department of Agriculture, Forest Service, In- termountain Research Station. 34p. Klopatek, J. M. 1987. Nitrogen Mineralization and Nitri- fication in Mineral Soils of Pinon-juniper Ecosystems. Soil Science Society of America Journal. 51:453-457. Klute, Arnold, ed. 1986. Methods of Soil Analysis - Part I, Physical and Mineralogical Methods. Monograph 9. Madison, WI: American Society of Agronomy. 1216. Kruse, William H. 1994. Effects of fuelwood harvesting on small mammal populations in a pinon-juniper woodland. P-J Symposium: Desired Future Condi- tions for P-J Ecosystems. 8-12 Aug. 1994 Flagstaff, AZ. Ludwig, John A.; Reynolds, James F. 1988. Statistical Ecology. New York: John Wiley and Sons. 337. Milliken, George A.; Johnson, Dallas E. 1984. Analysis of Messy Data Volume I: Designed Experiments. New York: Van Nostrand Reinhold Co. 473. Page, A. L.; Miller, R. H.; Keeney, D. R. eds. 1982. Meth- ods of Soil Analysis, Part 2 - Chemical and Microbi- ological Properties, 2nd ed. Agronomy, No. 9. Madi- son, WI. American Society of Agronomy. 1159. Smith, Robert L. 1966. Ecology and field biology. New York: Harper and Row, Inc. 686. Soil Survey Staff. 1975. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Sur- veys. Agric. Handb. 436. Washington, DC: U.S. Depart- ment of Agriculture, Soil Conservation Service. 754. 224 The Composition of Oils in Pinus edulis Michael Blair, Telletha Valenski, Andrew Sykes, Russell Balda, and Gerald Caple1 Abstract. — Pihon nut oil was reacted with sodium methoxide in methanol to yield the free methyl esters. These methyl esters were analyzed by GC- mass spectrometry. The pihon seeds are 60% by weight oil; of this oil 90% were unsaturated fatty acids with 41% being mono 9-octadecenoic acid and 59% being 9,12-octadecadienoic acid. Small amounts of palmitic acid and stearic acid, both saturated oils, were found, as were some C20 species. Preliminary studies indicated the actual oil constituents varied in pihon seeds from different areas. The highly unsaturated oil of Pinus edulis seeds might be a useful dietary supplement, with the current health concerns about the lowering of saturated fats in the diet. INTRODUCTION Foraging birds such as Pinon Jays, Scrub Jays, and Clark's Nutcrackers, instinctively harvest pinon nuts for the high supplement of fat. They discriminate between good and bad seeds, cones with a large number of good seeds, and trees that produced cones with a large number of good seeds (Vander Wall and Balda, 1977). The pihon nut provides the birds with good nourishment throughout the cold winter months. This can also be said for the Native people who also forage off the nut. Here we report on the composition of oils found in the pihon nut. The percentage of oil found in the pihon nut (Pinus edulis) is 58-62% (C.W. Bedkin, L.B. Shires). We have determined the ratios of saturated and unsaturated fats in pihon nuts by Gas chromotagraphy/Mass Spectroscopy (GC/MS) and compared these ratios to other com- mercial oils. The seeds were found from different areas of the southwest which gave variations in percentage of oil. In addition, a one seed analysis was done to make a comparison with larger quan- tities. MATERIALS AND METHODS After weighing the crushed nut, we separated the oil from the pihon nut by using a nonpolar 1 Department of Chemistry and Biology, Northern Arizona Univer- sity, Flagstaff, AZ. solvent (hexane) to extract the oil, and then filter- ing the solution. The powder extraction was set aside to do a protein and carbohydrate analysis. After evaporating the solvent, the oil was reweighed to give the percent oil in the seed, which agreed with the range given in the intro- duction. For the preparation of methyl esters from the oil extraction, triglycerides were reacted with Sodium Methoxide (Olsson, Urban; Kaufmann, Peter; Hersolof G. Bengt). Methyl esters were separated from the nut and characterized by GC/MS. Methyl linoleate, methyl oleate, methyl stearate, and methyl palmitate were purchased from Aldrich Chemical and used as standards. GC-Mass Spectroscopy spectra were collected on an HP 5890 Series II GC fitted with a 12m * 0.2mm * 0.33um HP-1 crosslinked methyl silicone gum column interfaced to an HP 5971A Mass Se- lective Detector. RESULTS/CONCLUSION The specific oils in Pinus edulis were character- ized as the methyl esters by GC/MS. This could be done on any quantity of pihon nuts, including a single nut analysis. Figure 1 is a typical GC/MS spectrum of a single nut analysis. We identified four major oils present in Pinus edulis plus a small percentage of higher C20 oils. The molecular weights and retention times of these methyl esters 225 Table 1 . — Retention Times of Methyl Esters (minutes). O a Imitate ^taa rata 11.1 Moi. wgt. 270 292 294 296 298 Qhollorl/Foht oneneu ^reDj 15.97 Shoulder 18.31 18.34 Shelled(June) 17.42 Shoulder 19.22 19.38 Freezer Nuts 17.1 18.55 18.77 18.85 19.05 Palmitate 270 17.19 Linolenate 292 18.82 Linoleate 294 18.92 Oleate 296 18.96 Stearate 298 19.04 are given in Table 1. The molecular weights and the retention times of methyl esters purchased from Aldrich closely match the methyl esters found in the pirion nut. The shelled nuts (Feb) were run under different GC conditions (higher Helium column pressure) accounting for the faster in re- tention times. The freezer nut sample obtained from a commercially available source and of uni- dentified history also contained an additional unsaturated oil at 18.55 min (asterisk-Figure 2). 3 o'd oob'c" T IC; 1 i ?aC000C ll 2600000 I : i 24QOOOC- 2200000 ■ •i ! ! <\ \\ 2000000- 1 SOOC'JO 1600C00- 1400COO - 1200000- 1000000- S OOCO j - I ooocoo- ft :i ii i 1 i i 200000- I] jj i ! i i ,i i n 0 i ili 1? 00 ll'.S'. ) IS ! 00 18. 50 , • | . i ■ '— i H"1 19. CO 19.50 20. CO 20.50 Figure 1. — Graph showing GC/MS characterization of methyl esters. Composition of Oil 0 20 40 SO 80 100 Figure 2. — Fat content of commercial oils including pihon. Figure 2 compares the fat content of Pinus edulis to other common commercial oils (USDA Handbook # 8). The graph shows the nut is high in unsatu- rated fat (90%), but has its own unique content of mono and diunsaturated fats. The exact ratio of the mono and diunsaturated esters were determined by digital transformation of the linoleate and oleate peaks and fitting the data to two Gaussian peaks using a commercially available computer program (Peak Fit). The relative amounts of linoleate to oleate esters are 59% to 41% by this technique. Integration using the GC/MS software incorrectly determined the relative percentages as 44/56% for the linoleate/oleate ratio, almost the opposite from above, due to poor resolution of these peaks. In the future we would like to develop a greater separation of the methyl esters using GC/MS. Preliminary evidence shows that a large number of isomers exist for methyl linoleate and methyl oleate. We would also like to know if seasonal changes cause the pinon nuts to have different percentage of oil. LITERATURE CITED Bader, Alfred; Harvey David; Nagarkatti, Jai. 1990. Aldrich Chemical Company, Inc. Bedkin, C. W; Shires, L. B. 1948. The Composition and Value of pinon Nuts. Agricultural Experimental Station. New Mexico of A &M State College, New Mexico. Christensen, Kerry M.; Whitham, Thomas G.; Balda, Russell E 1991. Discrimination among pinyon pine trees by Clark's Nutcrackers: effect of cone crop size and cone characters. Oecologia. 86: 402^07. Olsson, Urban; Kaufmann, Peter; Herslof, Bengt G. 1990. Multivariate optimization of a gas-liquid chroma- tographic analysis of fatty acid methyl esters of black currant seed oil. Journal of Chromatography 505: 385- 394. Watt, Bernice K. 1963. Composition of Foods. Agriculture Handbook No. 8. Agricultural Research Service. United States Department of Agriculture. 8: 135. •U.S. GOVERNMENT PRINTING OFFICE : 1995-674-451/25030 52S*LJ! 1022399924 The United States Department of Agriculture (USDA) prohibits discrimination in its programs on the basis of race, color, national origin, sex, religion, age, disability, political beliefs and marital or familial status. (Not all prohibited bases apply to all programs.) Persons with disabilities who require alternative means for communication of program information (braille, large print, audiotape, etc.) should contact the USDA Office of Communications at (202) 720-5881 (voice) or (202) 720-7808 (TDD). To file a complaint, write the Secretary of Agriculture, U.S. Department of Agriculture, Washington, D.C. 20250, or call (202) 720-7327 Voice) or (202) 720-1127 (TDD). USDA is an equal employment opportunity employer. NATIONAL AGRICULTURAL LIBRARY 1022399924 Rocky Mountains Southwest U.S. Department of Agriculture Forest Service Rocky Mountain Forest and Range Experiment Station The Rocky Mountain Station is one of eight regional experiment stations, plus the Fdrest Products Laboratory and the Washington Office Staff, that make up the Forest Service research organization. ■ RESEARCH FOCUS Research programs at the Rocky Mountain Station are coordinated with area universities and with other institutions. Many studies are conducted on a cooperative basis to accelerate solutions to problems involving range, water, wildlife and fish habitat, human and community development, timber, recreation, protection, and multiresource evaluation. RESEARCH LOCATIONS Research Work Units of the Rocky Mountain Station are operated in cooperation with universities in the following cities: ( Great Plains 'Station Headquarters: Albuquerque, New Mexico •Flagstaff, Arizona Fort Collins, Colorado* Laramie, Wyoming Lincoln, Nebraska Rapid City, South Dakota 240 W. Prospect Rd. Fort Collins, CO 80526