Academia.eduAcademia.edu
United States Department of Agriculture Central Appalachians Forest Ecosystem Vulnerability Assessment and Synthesis: A Report from the Central Appalachians Climate Change Response Framework Project Forest Service Northern Research Station General Technical Report NRS-146 February 2015 ABSTRACT Forest ecosystems across the Central Appalachians will be afected directly and indirectly by a changing climate over the 21st century. This assessment evaluates the vulnerability of nine forest ecosystems in the Central Appalachian Broadleaf Forest-Coniferous Forest-Meadow and Eastern Broadleaf Forest Provinces of Ohio, West Virginia, and Maryland for a range of future climates. We synthesized and summarized informaion on the contemporary landscape, provided informaion on past climate trends, and described a range of projected future climates. This informaion was used to parameterize and run muliple vegetaion impact models, which provided a range of potenial tree responses to climate. Finally, we brought these results before a mulidisciplinary panel of scienists, land managers, and academics familiar with the forests of this region to assess ecosystem vulnerability through a formal consensus-based expert elicitaion process. The summary of the contemporary landscape ideniies major forest trends and stressors currently threatening forests in the region. Observed trends in climate over the past century reveal that average minimum temperatures have increased in the area, paricularly in summer and fall. Precipitaion has also increased in the area, paricularly in fall. Projected climate trends for the next 100 years using downscaled global climate model data indicate a potenial increase in mean annual temperature of 2 to 8 °F for the assessment area. Projecions for precipitaion indicate increases in winter and spring precipitaion, and summer and fall precipitaion projecions vary by scenario. We ideniied potenial impacts on forests by incorporaing these future climate projecions into three forest impact models (DISTRIB, LINKAGES, and LANDIS PRO). Model projecions suggest that many mesic species, including American beech, eastern hemlock, eastern white pine, red spruce, and sugar maple may fare worse under future condiions, but other species such as eastern redcedar may beneit from projected changes in climate. Published literature on climate impacts related to wildire, invasive species, and forest pests and diseases also contributed to the overall determinaion of climate change vulnerability. We assessed vulnerability for nine forest ecosystems in the assessment area. The assessment was conducted through a formal elicitaion process of 19 science and management experts from across the area, who considered vulnerability in terms of the potenial impacts on a forest ecosystem and the adapive capacity of the ecosystem. Appalachian (hemlock)/northern hardwood forests, large stream loodplain and riparian forests, small stream riparian forests, and spruce/ir forests were determined to be the most vulnerable ecosystems. Dry/mesic oak forests and dry oak and oak/pine forests and woodlands were perceived as less vulnerable to projected changes in climate. These projected changes in climate and the associated impacts and vulnerabiliies will have important implicaions for economically valuable imber species, forestdependent wildlife and plants, recreaion, and long-term natural resource planning. Cover Photo Small stream riparian forest with a red spruce forest in the background. Photo by Patricia Butler, Northern Insitute of Applied Climate Science and Michigan Tech, used with permission. Manuscript received for publication June 2014 Published by: For additional copies, contact: USDA FOREST SERVICE 11 CAMPUS BLVD., SUITE 200 NEWTOWN SQUARE, PA 19073-3294 USDA Forest Service Publications Distribution 359 Main Road Delaware, OH 43015-8640 Fax: 740-368-0152 February 2015 Visit our homepage at: http://www.nrs.fs.fed.us/ Central Appalachians Forest Ecosystem Vulnerability Assessment and Synthesis: A Report from the Central Appalachians Climate Change Response Framework Project Patricia R. Butler, Louis R. Iverson, Frank R. Thompson III, Leslie A. Brandt, Stephen D. Handler, Maria K. Janowiak, P. Danielle Shannon, Chris Swanston, Kent Karriker, Jarel Barig, Stephanie Connolly, William D. Dijak, Scot Bearer, Steve L. Blat, Andrea Brandon, Elizabeth Byers, Cheryl Coon, Tim Culbreth, Jad Daly, Wade Dorsey, David Ede, Chris Euler, Neil Gillies, David M. Hix, Catherine Johnson, Latasha Lyte, Stephen Mathews, Dawn McCarthy, Dave Minney, Daniel Murphy, Claire O’Dea, Rachel Orwan, Mathew Peters, Anantha Prasad, Coton Randall, Jason Reed, Cynthia Sandeno, Tom Schuler, Lesley Sneddon, Bill Stanley, Al Steele, Susan Stout, Randy Swaty, Jason Teets, Tim Tomon, Jim Vanderhorst, John Whatley, and Nicholas Zegre AuThoRS PATRICIA R. BUTLER is a climate change outreach specialist with the Northern Insitute of Applied Climate Science, Michigan Technological University, U.J. Noblet Forestry Building, 1400 Townsend Drive, Houghton, MI 49931, prbutler@mtu.edu. LOUIS R. IVERSON is a landscape ecologist with the U.S. Forest Service, Northern Research Staion, 359 Main Road, Delaware, OH 43015, liverson@fs.fed.us. FRANK R. THOMPSON III is a research wildlife biologist with the U.S. Forest Service, Northern Research Staion, 202 AnheuserBusch Natural Resources Building, University of Missouri – Columbia, Columbia, MO 65211, frthompson@fs.fed.us. LESLIE A. BRANDT is a climate change specialist with the Northern Insitute of Applied Climate Science, U.S. Forest Service, Northern Research Staion, 1992 Folwell Avenue, St. Paul, MN 55108, lbrandt@fs.fed.us. STEPHEN D. HANDLER is a climate change specialist with the Northern Insitute of Applied Climate Science, U.S. Forest Service, 410 MacInnes Drive, Houghton, MI 49931, sdhandler@fs.fed.us. MARIA K. JANOWIAK is a climate change adaptaion and carbon management scienist with the Northern Insitute of Applied Climate Science, U.S. Forest Service, 410 MacInnes Drive, Houghton, MI 49931, mjanowiak02@fs.fed.us. P. DANIELLE SHANNON is a climate change outreach specialist with the Northern Insitute of Applied Climate Science, Michigan Technological University, U.J. Noblet Forestry Building, 1400 Townsend Drive, Houghton, MI 49931, dshannon@mtu.edu. CHRIS SWANSTON is a research ecologist with the U.S. Forest Service, Northern Research Staion, and director of the Northern Insitute of Applied Climate Science, 410 MacInnes Drive, Houghton, MI 49931, cswanston@fs.fed.us. KENT KARRIKER is the ecosystems group leader with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241, kkarriker@fs.fed.us. JAREL BARTIG is an ecologist with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, jbarig@fs.fed.us. STEPHANIE CONNOLLY is a forest soil scienist with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241, sconnolly@fs.fed.us. WILLIAM D. DIJAK is a wildlife biologist and geographic informaion systems specialist with the U.S. Forest Service, Northern Research Staion, 202 Anheuser-Busch Natural Resources Building, University of Missouri – Columbia, Columbia, MO 65211, wdijak@fs.fed.us. SCOTT BEARER is a senior conservaion scienist with The Nature Conservancy, 220 West Fourth Street, 3rd Floor, Williamsport, PA 17701, sbearer@tnc.org. STEVE L. BLATT is a forest biologist with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, sblat@fs.fed.us. ANDREA BRANDON is the Central Appalachians program coordinator with The Nature Conservancy, 194 Airport Road, Elkins, WV 26241, abrandon@tnc.org. ELIZABETH BYERS is a vegetaion ecologist with the West Virginia Division of Natural Resources, Natural Heritage Program, P.O. Box 67, Elkins, WV 26241, elizabeth.a.byers@wv.gov. CHERYL COON is a forest botanist with the Hoosier Naional Forest, 811 Consituion Avenue, Bedford, IN 47421, ccoon@fs.fed.us. TIM CULBRETH is the Chesapeake Watershed Forester for the Maryland Forest Service, 580 Taylor Avenue, E-1, Annapolis, MD 21401, tculbreth@dnr.state.md.us. JAD DALEY is a program director at the Trust for Public Land, 660 Pennsylvania Avenue SE, Suite 401, Washington, DC 20003, jad.daley@tpl.org. WADE DORSEY is the forest manager of the Savage River State Forest, 127 Headquarters Lane, Grantsville, MD 21536, wdorsey@dnr.state.md.us. DAVID EDE (reired) was the forest planner and Naional Environmental Policy Act (NEPA) coordinator with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241. He can be reached at ededm86@gmail.com. CHRIS EULER is an archaeologist with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, ceuler@fs.fed.us. NEIL GILLIES is the director of science and educaion at the Cacapon Insitute, P.O. Box 68, High View, WV, 26808, ngillies@cacaponinsitute.org. DAVID M. HIX is an associate professor with the School of Environment and Natural Resources at Ohio State University, 365A Kotman Hall, 2021 Cofey Road, Columbus, OH 43210, hix.6@osu.edu. CATHERINE JOHNSON is a wildlife biologist with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241, catherinejohnson@fs.fed.us. LATASHA LYTE is a forest soil scienist with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, ljlyte@fs.fed.us. STEPHEN MATTHEWS is a research assistant professor with the School of Environment and Natural Resources at Ohio State University, and an ecologist with the U.S. Forest Service, Northern Research Staion, 359 Main Road, Delaware, OH 43015, snmathews@fs.fed.us. DAWN McCARTHY is the assistant district ranger for operaions with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, dmccarthy02@fs.fed.us. DAVE MINNEY is a ire ecologist with Allegheny Ecological Services, 2315 Wickline Road, Beaver, OH 45613, dminney57@gmail.com. DANIEL MURPHY is an assistant professor of anthropology at the University of Cincinnai, 462 Braunstein Hall, Cincinnai, OH 45221, murphdl@ucmail.uc.edu. CLAIRE O’DEA is an air quality specialist with the U.S. Forest Service, Eastern Region, 400 Independence Avenue, SW., Mailstop: 1121, Washington, DC 20250, cbodea@fs.fed.us. RACHEL ORWAN is the NEPA, objecions and liigaion coordinator with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, rorwan@fs.fed.us. MATTHEW PETERS is a geographic informaion systems technician with the U.S. Forest Service, Northern Research Staion, 359 Main Road, Delaware, OH 43015, mathewpeters@fs.fed.us. ANANTHA PRASAD is an ecologist with the U.S. Forest Service, Northern Research Staion, 359 Main Road, Delaware, OH 43015, aprasad@fs.fed.us. COTTON RANDALL is the special projects administrator with the Ohio Division of Forestry, 2045 Morse Road, H-1, Columbus, OH 43229, coton.randall@dnr.state.oh.us. JASON REED is the engineering and natural resources group leader with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241, jareed02@fs.fed.us. CYNTHIA SANDENO is an ecologist with the Monongahela Naional Forest, 200 Sycamore Street, Elkins, WV 26241, cmsandeno@fs.fed.us. TOM SCHULER is a project leader and research forester with the Fernow Experimental Forest, P.O. Box 404, Parsons, WV 26287, tschuler@fs.fed.us. LESLEY SNEDDON is a naional ecologist with NatureServe, c/o University of Massachusets at Boston, Biology Department, 100 Morissey Boulevard, Boston, MA 02125, lesley_sneddon@natureserve.org. BILL STANLEY is the director of conservaion with The Nature Conservancy in Ohio, 6375 Riverside Drive, Suite 100, Dublin, OH 43017, bstanley@tnc.org. AL STEELE is a physical scienist with the U.S. Forest Service, Northeastern Area State & Private Forestry, 180 Canield Street, Morgantown, WV 26505, asteele@fs.fed.us. SUSAN STOUT is a project leader with the U.S. Forest Service, Northern Research Staion, 359 Main Road, Delaware, OH 43015, sstout@fs.fed.us. RANDY SWATY is an ecologist with The Nature Conservancy, 101 South Front Street, Suite 105, Marquete, MI 49855, rswaty@tnc.org. JASON TEETS is an ecological site descripion specialist with the Natural Resources Conservaion Service, 201 Scot Avenue, Morgantown, WV 26508, jason.teets@wv.usda.gov. TIM TOMON is a forest entomologist with the West Virginia Department of Agriculture, Plant Industries Division, 1900 Kanawha Boulevard, E., Charleston, WV 25305, tomon@wvda.us. JIM VANDERHORST is a vegetaion ecologist with the West Virginia Division of Natural Resources, Natural Heritage Program, P.O. Box 67, Elkins, WV 26241, james.p.vanderhorst@wv.gov. JOHN WHATLEY is a zone ire management oicer with the Wayne Naional Forest, 13700 U.S. Hwy 33, Nelsonville, OH 45764, jwhatley@fs.fed.us. NICHOLAS ZEGRE is an assistant professor of hydrology at West Virginia University, Division of Forestry and Natural Resources, P.O. Box 6125, Morgantown, WV 26506, nicolas.zegre@mail.wvu.edu. Page intentionally left blank PREFACE CoNTEXT AND SCoPE This assessment is a fundamental component of the Central Appalachians Climate Change Response Framework project. The Framework is a collaborative, cross-boundary approach among scientists, managers, and landowners to incorporate climate change considerations into natural resource management. Six Framework projects are currently underway, covering approximately 250 million acres in the northeastern and midwestern United States: Northwoods, Central Appalachians, Central Hardwoods, Mid-Atlantic, New England, and Urban. Each project interweaves four components: science and management partnerships, vulnerability assessments, adaptation resources, and demonstration projects. We designed this assessment to be a synthesis of the best available scientific information on climate change and forest ecosystems. Its primary goal is to inform forest managers in the Central Appalachians region, in addition to people who study, recreate, and live in these forests. As new scientific information arises, our understanding of climate change and forest ecosystems will be strengthened. Most importantly, this assessment does not make recommendations about how this information should be used. The scope of the assessment is terrestrial forest ecosystems, with a particular focus on tree species. Climate change will also have impacts on aquatic systems, wildlife, and human systems, but addressing these issues in depth is beyond the scope of this assessment. The large list of authors reflects the highly collaborative nature of this assessment. The overall document structure and much of the language was a coordinated effort among Leslie Brandt, Patricia Butler, Maria Janowiak, Stephen Handler, and Chris Swanston. Danielle Shannon conducted much of the data analysis and developed maps for Chapters 1, 3, and 4. Louis Iverson, Stephen Matthews, Matthew Peters, and Anantha Prasad provided and interpreted Tree Atlas information for Chapter 5, and assisted with the data processing for the climate data presented in Chapter 4. Frank Thompson and William Dijak provided results and interpretation of the LINKAGES and LANDIS PRO models. All modeling teams coordinated their efforts impressively. Kent Karriker, Jarel Bartig, and Stephanie Connolly provided substantial input throughout the document. Among the many others who made valuable contributions to the assessment, Scott Pugh (U.S. Department of Agriculture, Forest Service, Forest Inventory and Analysis [FIA] Program) provided technical and analytical support for querying FIA databases. We also thank Kevin Potter (North Carolina State University), James Rentch (West Virginia University), and an additional reviewer, who provided formal technical reviews of the assessment. Their thorough reviews greatly improved the quality of this assessment. Page intentionally left blank CoNTENTS Execuive Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1 Introducion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8 Chapter 1: The Contemporary Landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 Chapter 2: Climate Change Science and Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Chapter 3: Observed Climate Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 Chapter 4: Projected Changes in Climate, Extremes, and Physical Processes . . . . . . . . . . . . . . . . . 89 Chapter 5: Future Climate Change Impacts on Forests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Chapter 6: Forest Ecosystem Vulnerabiliies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 Chapter 7: Management Implicaions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192 Literature Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199 Appendix 1: Species Lists . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234 Appendix 2: Trend Analysis and Historical Climate Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238 Appendix 3: Addiional Future Climate Projecions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243 Appendix 4: Addiional Impact Model Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254 Appendix 5: Vulnerability and Conidence Determinaion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Page intentionally left blank EXECuTiVE SuMMARY This assessment evaluates key vulnerabilities for forest ecosystems in the Central Appalachian Broadleaf Forest-Coniferous Forest-Meadow and Eastern Broadleaf Forest Provinces of Ohio, West Virginia, and Maryland across a range of future climate scenarios. This assessment was completed as part of the Central Appalachians Climate Change Response Framework project, a collaborative approach among researchers, managers, and landowners to incorporate climate change considerations into forest management. ecosystems, and how climate may interact with other stressors on the landscape. The assessment summarizes current conditions and key stressors and identifies past and projected trends in climate. This information is then incorporated into model projections of future forest change. These projections, along with published research and local knowledge and expertise, are used to identify the factors that contribute to the vulnerability of nine major forest ecosystems within the assessment area through the end of this century. A final chapter summarizes the implications of these impacts and vulnerabilities on a variety of forest-related ecological, social, and economic topics across the region. • Fragmentation and land-use change • Shifts in natural disturbance regimes (e.g., shifts in drought or flood frequencies) • Forest diseases and insect pests • Nonnative plant species invasion • Shifts in fire regime • Mineral, gas, and wind energy development • Erosion and sedimentation Main Points ● The assessment area of the Central Appalachians region contains about 29 million acres, of which 18.9 million acres are forest land. Private individuals and organizations own more than 85 percent of forest land. ● Current major stressors and threats to forest ecosystems in the region are: ● Repeated periods of warming and cooling over the last 15,000 years have resulted in multiple waves of species retracting and expanding from the south and from climatic refuges along the Atlantic coast. ● Historical land use and past management practices (17th century onward) have resulted in second-growth forests that have been rebounding from large-scale deforestation and wildfire. Secondary forests are largely even-aged with poor structure and reduced species diversity. ● The forest products and forest-related recreation industries are major contributors to the region’s economy, and an increasing amount of the forest land in the assessment area is managed according to at least one sustainability certification standard. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE This chapter describes the forests and related ecosystems across the Central Appalachian Broadleaf Forest-Coniferous Forest-Meadow and Eastern Broadleaf Forest Provinces of Ohio, West Virginia, and Maryland and summarizes current threats and management trends. This information lays the foundation for understanding how shifts in climate may contribute to changes in forest 1 EXECuTiVE SuMMARY ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG ChAPTER 3: oBSERVED CLiMATE ChANGE This chapter provides a brief background on climate change science, models that simulate future climate change, and models that project the effects of climate change on tree species and ecosystems. This chapter also describes the climate data used in this assessment. Many of the climatic changes that have been observed across the world over the past century are also evident in the assessment area. This chapter summarizes our current understanding of observed changes and current climate trends in the assessment area and across the Central Appalachians region, with a focus on the last 100 years. Main Points ● Temperatures have been increasing at a global scale and across the United States over the past century. ● Climate scientists attribute this increase in temperature to human activities. ● Major contributors to warming are greenhouse gases from fossil fuel burning, agriculture, and changes in land use. Main Points ● Annual minimum temperatures increased over the past century, with summer and fall minimum temperatures warming the most rapidly. April, June, July, August, and November had the greatest increases in minimum temperature. Maximum temperatures decreased during July, September, and October. Hot days are occurring more frequently. Hemlock on the Monongahela Naional Forest, West Virginia. Photo by Patricia Butler, Northern Insitute of Applied Climate Science (NIACS) and Michigan Tech, used with permission. 2 EXECuTiVE SuMMARY ● ● ● Precipitation patterns have changed across the region, with the most change occurring in fall (increase of 2.3 inches). The number of intense precipitation events has increased. Snowfall decreased across the assessment area, and lake ice duration has declined. Climate change is also indicated by positive trends in growing season length, shifts in flowering phenology, and changes in wildlife emergence and migration. ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES This chapter describes climate projections for the assessment area over the 21st century, including projections related to patterns of extreme weather events and other climate-related processes. Temperature and precipitation projections are derived from downscaled simulations of climate models. Published scientific literature provides the basis for describing possible trends in a range of climate-driven processes, such as extreme weather events and snowfall. ● The number of hot days is expected to increase and the number of cold days is projected to decrease. ● Intense precipitation events are expected to become more frequent. ● Streamflow and flooding potential are expected to increase in the winter and spring, and decrease in the summer and fall. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS This chapter summarizes the potential impacts of climate change on forests in the assessment area, drawing on information from a coordinated series of model simulations and published research. Main Points ● Many temperate tree species present within the assessment area are expected to tolerate a mild degree of warming, but are expected to decline under higher rates of warming. ● Many mesic species, including American beech, eastern hemlock, eastern white pine, red spruce, and sugar maple are among those projected to have reductions in suitable habitat, growth potential, and biomass under a high degree of warming over the next century. ● Many species are expected to lose establishment and regeneration potential over the next century, but in the absence of other mortality factors, may persist as mature individuals that continue to grow for much longer. ● Species with ranges that extend largely to the south such as eastern redcedar, post oak, and shortleaf pine may have increases in suitable habitat and biomass. Loblolly pine, currently only in plantations in the assessment area, is also expected to fare well under the future climate. Main Points ● ● Temperatures are expected to increase over the next century, under a range of climate scenarios and in all seasons. Precipitation is projected to increase in winter and spring across a range of climate scenarios. Projections of summer and fall precipitation are more variable; depending on the scenario, precipitation is projected to decrease during either summer or fall. ● Late season droughts or localized soil moisture deficits are expected to become more frequent. ● The growing season length is expected to increase by up to a month. 3 EXECuTiVE SuMMARY ● The model projections used in this assessment do not account for many other factors that may change under a changing climate. Scientific literature was used to provide additional information on these factors, including: • Drought stress • Wildfire frequency and severity • Acid deposition and carbon dioxide fertilization • Altered nutrient cycling • Changes in invasive species, insect pests, and forest diseases • Effects of herbivory on young regeneration • Interactions among these factors ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES change, with an emphasis on shifts in dominant species, system drivers, and stressors. The adaptive capacity of forest systems was also examined as a key component to overall vulnerability. Synthesis statements are provided to capture general trends. Detailed vulnerability determinations are also provided for nine forest ecosystems (Table 1). We consider a system to be vulnerable if it is at risk of a composition change leading to a new identity, or if the system is anticipated to suffer substantial declines in acreage, health, or productivity. Main Points Potenial Impacts of Climate Change on Drivers and Stressors ● This chapter focuses on the vulnerability of major forest ecosystems in the assessment area to climate Temperatures will increase (robust evidence, high agreement). All downscaled climate models project that average temperatures will increase across much of the assessment area. Table 1.—Vulnerability determinaion summaries for forest ecosystems considered in this assessment Potenial impacts Adapive capacity Vulnerability Evidence Agreement Negaive Low-Moderate High Medium Medium Dry calcareous forest, woodland, and glade Neutral-Negaive Low-Moderate Moderate-High Limited-Medium Medium Dry oak and oak/pine forest and woodland Posiive Moderate-High Low Medium Medium-High Dry/mesic oak forest Posiive-Neutral High Low- Moderate Medium Medium-High Large stream loodplain and riparian forest Negaive Low High Medium Medium Mixed mesophyic and cove forest Neutral-Negaive Moderate-High Moderate Limited-Medium Medium North-central interior beech/maple forest Neutral Moderate Moderate Limited-Medium Medium Small stream riparian forest Negaive Moderate Moderate-High Medium Medium Spruce/ir forest Negaive Moderate High Limited-Medium Medium Forest ecosystem Appalachian (hemlock)/ northern hardwood forest 4 EXECuTiVE SuMMARY ● ● ● Growing seasons will get longer (robust evidence, high agreement). There is high agreement among evidence that projected temperature increases will continue the current trend of longer growing seasons in the assessment area. The amount and timing of precipitation will change (medium evidence, high agreement). All downscaled climate models agree that there will be changes in precipitation patterns across the assessment area. Intense precipitation events will continue to become more frequent (medium evidence, medium agreement). There is some agreement that the number of heavy precipitation events will continue to increase in the assessment area. If so, impacts from flooding and soil erosion may also become more damaging. will lead to increases in certain pest and pathogen outbreaks, but research to date has examined few species in the assessment area. ● Many invasive plants will increase in extent or abundance (medium evidence, high agreement). Evidence indicates that an increase in temperature and more frequent disturbances will lead to increases in many invasive plant species. Potenial Impacts of Climate Change on Forests ● Suitable habitat for northern species will decline (medium evidence, high agreement). All three impact models project a decrease in suitability for northern species such as eastern hemlock, red spruce, and sugar maple, compared to current climate conditions. ● Severe storms will increase in frequency and severity (medium evidence, medium agreement). There is some agreement that future climate change will destabilize atmospheric circulation patterns and processes, leading to increased risk of severe weather. ● Habitat is projected to become more suitable for southern species (medium evidence, high agreement). All three impact models project an increase in suitability for southern species such as eastern redcedar and loblolly pine, compared to current climate conditions. ● Soil moisture patterns will change (medium evidence, high agreement), with drier soil conditions later in the growing season (medium evidence, medium agreement). Studies show that climate change will have impacts on soil moisture, but there is some disagreement among climate and impact models on how soil moisture will change during the growing season. ● Species composition will change across the landscape (limited evidence, high agreement). Although few models have specifically examined how species composition may change, model results from individual species, paleoecological data, and ecological principles suggest that recognized communities may dissolve to form new mixes of species. ● ● Climate conditions will increase wildfire risk by the end of the century (medium evidence, medium agreement). Some national and global studies suggest that wildfire risk will increase in the region, but few studies have specifically looked at wildfire potential in the assessment area. ● Certain insect pests and pathogens will increase in occurrence or become more damaging (medium evidence, high agreement). Evidence indicates that an increase in temperature A major transition in forest composition is not expected until after the middle of the century (2040 to 2069) (medium evidence, medium agreement). Although some models indicate major changes in habitat suitability, results from spatially dynamic forest landscape models indicate that a major shift in forest composition across the landscape may take 100 years or more in the absence of major disturbances. 5 EXECuTiVE SuMMARY ● Climate change is expected to affect early growth and regeneration conditions (medium evidence, medium agreement). Seedlings are more vulnerable than mature trees to changes in temperature, moisture, and other seedbed and early growth requirements. ● Increased fire frequency and harvesting will accelerate shifts in forest composition across the landscape (medium evidence, medium agreement). Studies from other regions show that increased fire frequency can accelerate the decline of species negatively affected by climate change and can accelerate the northward migration of southern tree species. ● Net change in forest productivity is expected to be minimal (medium evidence, low agreement). A few studies have examined the impact of climate change on forest productivity, but they disagree on how multiple factors may interact to influence it. Adapive Capacity Factors ● Low-diversity ecosystems are at greater risk (medium evidence, high agreement). Studies have consistently shown that diverse ecosystems are more resilient to disturbance, and lowdiversity ecosystems are more vulnerable to change. ● Species in fragmented landscapes will have less opportunity to migrate long distances in response to climate change (limited evidence, high agreement). Evidence suggests that species may not be able to disperse overthe distances required to keep up with climate change, but little research has been done in the region on this topic. ● Ecosystems that are highly limited by hydrologic regime or geological features may be topographically constrained (limited evidence, medium agreement). Our current understanding of the ecology of Central Appalachians ecosystems suggests that some species will be unable to migrate to new areas due to topographic constraints.  ● Ecosystems that are tolerant of disturbance or are disturbance-adapted have less risk of declining on the landscape (medium evidence, high agreement). Basic ecological theory and other evidence support the idea that systems that are adapted to more frequent disturbance will be at lower risk. ● Fire-adapted ecosystems will be more resilient to climate change (high evidence, medium agreement). Studies have shown that fireadapted ecosystems are better able to recover after disturbances and can promote many of the species that are expected to do well under a changing climate. ● Ecosystems occupying habitat in areas of high landscape complexity have more opportunities for persistence in pockets of refugia (medium evidence, medium agreement). The diversity of landscape positions occupied by forest may provide opportunities for natural refugia, for example where cool air and moisture accumulate in valley bottoms. ChAPTER 7: MANAGEMENT iMPLiCATioNS This chapter summarizes the implications of potential climate change impacts on important facets of forest management and planning in the Central Appalachians region, such as impacts on wildlife or cultural resources. The process we used to assess the vulnerability of forest ecosystems was not used to consider these topics. Rather, we point out important implications, ongoing research, and sources for more information on how climate change is expected to affect these topics. This chapter does not make recommendations as to how management should be adjusted to cope with these impacts, because impacts and responses will vary by ecosystem, ownership, and management objective. EXECuTiVE SuMMARY Main Points ● Management of endemic plants and animals that depend on forests may face additional challenges as the climate shifts. ● Prevention and eradication of nonnative invasive plant species are expected to become more difficult and require more resources. ● The timing of activities, including prescribed fire, recreation, or timber removal may need to be shifted as temperatures and precipitation patterns change. ● Responses to increased risk of wildfire or large-scale wind and storm events may require reassessing emergency response plans, water resource infrastructure, and available resources. ● Climate change may present opportunities for the forest products industry, recreation, and other sectors if resource managers are able to anticipate and respond to changing conditions. A riparian hemlock community in West Virginia. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 7 iNTRoDuCTioN CoNTEXT This assessment is part of a regional effort called the Central Appalachians Climate Change Response Framework (Framework; www.forestadaptation. org). The Framework project was initiated in 2009 in northern Wisconsin with the overarching goal of helping managers incorporate climate change considerations into forest management. To meet the challenges brought about by climate change, a team of federal and state land management agencies, private forest owners, conservation organizations, and others have come together to accomplish three objectives: 1. Provide a forum for people working across the Central Appalachians to effectively and efficiently share experiences and lessons learned. 2. Develop new user-friendly information and tools to help land managers factor climate change considerations into decisionmaking. 3. Support efforts to implement actions for addressing climate change impacts in the Central Appalachians. The Framework process is designed to work at multiple scales. The Central Appalachians Framework is coordinated across the region, but activities are generally conducted at the state or local level to allow for greater specificity. Additionally, regional Framework projects are underway in several other regions: Central Hardwoods, MidAtlantic, New England, Northwoods, and an Urban pilot project in Chicago. 8 The Central Appalachians Framework is an expansion of the original northern Wisconsin effort, and has been supported in large part by the U.S. Forest Service. Across the Central Appalachians region, the project is being guided by an array of partners with an interest in forest management, including: • Northern Institute of Applied Climate Science • U.S. Forest Service, Eastern Region • U.S. Forest Service, Northern Research Station • U.S. Forest Service, Northeastern Area State & Private Forestry • Trust for Public Land • The Nature Conservancy • NatureServe • Natural Resources Conservation Service • Ohio Department of Natural Resources • West Virginia Division of Natural Resources • Maryland Department of Natural Resources This assessment is designed to provide detailed information for forest ecosystems across the Central Appalachians region. Several independent efforts related to climate change, natural ecosystems, and human well-being are also occurring at the state level. This assessment complements other assessments that have been created for the assessment area and for the broader Central Appalachians region. The Framework project will also work to integrate the results and outcomes from other projects related to climate change and natural resource management. iNTRoDuCTioN This assessment bears some similarity to other synthesis documents about climate change science, such as the National Climate Assessment (Melillo et al. 2014) and the Intergovernmental Panel on Climate Change (IPCC) reports (working group contributions to the Fifth Assessment at http://www. ipcc.ch/report/ar5/). Where appropriate, we refer to these larger-scale documents when discussing national and global changes. However, this assessment differs from these reports in many ways. This assessment was not commissioned by any federal government agency nor does it give advice or recommendations to any federal government agency. It also does not evaluate policy options or provide input into federal priorities. Instead, this report was developed by the authors to fulill a joint need of understanding local impacts of climate change on forests and assessing which tree species and forest ecosystems may be the most vulnerable in the Central Appalachians region. Although it was written to be a resource for forest managers, it is irst and foremost a scientiic document that represents the views of the authors. SCoPE AND GoALS The primary goal of this assessment is to summarize potential changes to the forest ecosystems of the Central Appalachians region under a range of possible future climates, and determine the vulnerability of forest ecosystems to these changes during the next century. Included is a synthesis of information about the current landscape as well as projections of climate and vegetation changes used to assess these vulnerabilities. Uncertainties and gaps in understanding are discussed throughout the document. This assessment covers about 18.9 million acres of forest land in Ohio, West Virginia, and Maryland (Fig. 1). The assessment area boundaries are deined by the Eastern Broadleaf Forest (Ecological Province 221) and the Central Appalachian Broadleaf ForestConiferous Forest-Meadow (Ecological Province M221) (McNab and Avers 1994, McNab et al. 2007). In addition to these state and ecological boundaries, we used county-level information that most closely Figure 1.—The assessment area overlaps two secions of the Eastern Broadleaf Forest Province (green) and three secions of the Central Appalachian Broadleaf Forest-Coniferous Forest-Meadow Province (blue) within Ohio, West Virginia, and Maryland (Cleland et al. 2007a). 9 iNTRoDuCTioN represented the assessment area when ecoregional data were not available, limiting our selections to the counties that are most analogous to the assessment area (21 Ohio counties, all West Virginia counties, and 3 Maryland counties). Chapter 3: Observed Climate Change provides information on the past and current climate of the assessment area, summarized from the interactive ClimateWizard database and published literature. This chapter also summarizes some relevant ecological indicators of observed climate change. Land ownership is fairly similar across the three states. Overall, more than 85 percent of forest land in the assessment area is owned by private individuals and organizations. Approximately 8 percent of land is federally owned, with the Wayne and Monongahela National Forests administering the bulk of federal lands. State agencies own 5 percent of forest land; and county, municipal, and local governments own 1.4 percent. This assessment synthesizes information covering all forest lands in the assessment area in recognition of the area’s dispersed patterns of forest composition and land ownership. Chapter 4: Projected Changes in Climate, Extremes, and Physical Processes presents downscaled climate change projections for the assessment area, including future temperature and precipitation data. It also includes summaries of other climate-related trends that have been projected within the assessment area and the broader Midwest and Northeast. ASSESSMENT ChAPTERS This assessment contains the following chapters: Chapter 1: The Contemporary Landscape describes existing conditions, providing background on the physical environment, ecological character, and broad socioeconomic dimensions of the assessment area. It defines the nine forest ecosystems we refer to in later chapters. Chapter 2: Climate Change Science and Modeling contains background on climate change science, projection models, and impact models. It also describes the techniques used in developing climate projections to provide context for the model results presented in later chapters. 10 Chapter 5: Future Climate Change Impacts on Forests summarizes ecosystem model results that were prepared for this assessment. Three modeling approaches were used to simulate climate change impacts on forests: a species distribution model (DISTRIB of the Climate Change Tree Atlas), and two forest simulation models (LINKAGES and LANDIS PRO). This chapter also includes a literature review of other climate-related impacts on forests that the models did not consider. Chapter 6: Forest Ecosystem Vulnerabilities synthesizes the potential effects of climate change on the forest ecosystems of the assessment area and provides detailed vulnerability determinations for nine major forest ecosystems. Chapter 7: Management Implications draws connections from the forest ecosystem vulnerability determinations to a wider network of related concerns shared by forest managers, including forest management, recreation, cultural resources, and forest-dependent wildlife. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE The Central Appalachians region is home to some of the most biologically diverse forests in North America. The diverse forests of the Central Appalachians provide many environmental, cultural, and economic benefits. This forested landscape sustains the people of the region by providing economically important forest products, outdoor recreation opportunities, and other services. This chapter includes a brief introduction to the complex factors that shape the forests in the region and provides context for the modeling results and interpretations provided in later chapters. LANDSCAPE SETTiNG The assessment area covers nearly 29 million acres, and is defined by a combination of ecological and political boundaries. The area is bounded by Ecological Provinces M221 (Central Appalachian Broadleaf Forest-Coniferous Forest-Meadow) and 221 (Eastern Broadleaf Forest) of the National Hierarchical Framework of Ecological Units, and by the state boundaries of Ohio, West Virginia, and Maryland (McNab et al. 2007). Provinces are broad geographic areas that share similar coarse features, such as climate, glacial history, and vegetation types. Provinces are divided by sections that are based on similarities in geologic parent material, elevation, plant distribution, and regional climate (McNab et al. 2007). To gain a better understanding of differences in forest ecosystems across the landscape, we focused on the five sections within these two provinces (Fig. 1). The major physical and biological features of the sections are summarized below. Physical Environment Climate The existing climates within the Central Appalachians are strongly influenced by atmospheric circulation patterns, latitude, topography, and abrupt changes in elevation. The primary factors influencing the climate are latitude and proximity to Lake Erie in the glaciated and gently dissected northern and western sections, and elevation and complex topography in the mountainous eastern sections. Three major air masses move through the assessment area. Hot, dry air from the southwest and cold, dry air from the north affect much of Ohio and West Virginia. Warm, moist air from the Gulf of Mexico sweeps east of the Allegheny Mountains to affect the eastern panhandle of West Virginia and Maryland (U.S. Geological Survey [USGS] 1989). Occasional easterlies can also sweep moist air from the Atlantic Ocean across Maryland to the Allegheny Mountains. The Allegheny Mountains (M221B) and Western Glaciated Allegheny Plateau (221F) have cooler and wetter climates than the rest of the assessment area, although microclimates within these sections are highly variable due to the effects of topography and relief. Average annual temperatures are 49 °F, with winters averaging 28 to 30 °F and winter minimum temperatures averaging 20 °F. Summers are also cooler in these sections, averaging 7 to 9 °F, with average maximum temperatures of 78 to 81 °F. In the highest elevations of M221B, daily minimum temperatures are the most extreme in the region, 11 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE and can reach -15 to -20 °F (U.S. Department of Agriculture 2012). The freeze-free growing season is less than 150 days, and can be shorter than 100 days in valleys that are subject to frost pocket effects (Koss et al. 1988). Annual precipitation may reach 70 inches in the high-elevation and lake-effect areas, whereas low-elevation and inland areas may get as little as 30 inches (National Oceanic and Atmospheric Administration [NOAA] 2014c). Annual snowfall is included in these averages and follows a similar pattern, with the high-elevation and lake-effect areas averaging more than 72 inches and lower elevation and inland areas receiving as little as 24 inches (NOAA 2014c). The warmest parts of the assessment area are Sections 221E, M221A, and M221C, where average annual temperatures range from 51 to 52 °F (Appendix 2). Winter average temperatures range from 32 to 34 °F, and minimum temperatures range from 22 to 24 °F. Summers are relatively hot, with average temperatures ranging from 70 to 71 °F, and maximum summer temperatures ranging from 82 to 83 °F. Precipitation is even more variable across the assessment area. In general, precipitation ranges from 35 inches per year in the lower elevation areas to 5 inches per year in the highest elevations (Chapter 3). The northernmost corner of Ohio’s Western Glaciated Allegheny Plateau receives slightly more precipitation because of its proximity to Lake Erie. Precipitation is also slightly higher on the western slopes of the Allegheny Mountains, where it can reach 70 inches per year in the highest elevations. Prevailing winds moving across Ohio and West Virginia pick up moisture, and release it as the air is forced to rise rapidly over the mountains. Moisture-laden air is obstructed by the mountain ridge, which allows precipitation and runoff to enter the western watershed, but not the eastern watershed, resulting in a rainshadow effect. Lake-effect snow from Lake Erie can also 12 generate winter storms, which are more frequent in northeastern Ohio, producing up to three snowstorms per decade with more than  inches of snow, and up to five ice storms per decade (Kunkel et al. 2013a). The frequency of these winter storms decreases to the south, with southeastern Ohio receiving an average of one -inch snowstorm and three ice storms per decade. Lake effect may combine with local weather processes to generate up to 30 percent of annual snowfall in the northern mountains of West Virginia (Kunkel et al. 2009a, 2009b). Extreme weather events in the area include highintensity rains associated with occasional hurricanes, short and infrequent drought periods, heat waves, windstorms and tornadoes, and ice storms (McNab et al. 2007). A more detailed description of past and contemporary climate of the region can be found in Chapter 3. Geology, Landform, and Soils The assessment area comprises both glaciated lands in northern Ohio and unglaciated, elevated lands in Ohio, West Virginia, and Maryland. There is much variation in elevation, from 17 feet in the Maryland portion of the assessment area to 4,81 feet at Spruce Knob in West Virginia (Fig. 2). Six groups of parent material formed the soils that dominate this assessment area. One group, residuum, developed in place by the weathering of underlying bedrock. Another group, colluvium, weathered from bedrock and was deposited at the base of steep slopes. Alluvium, lacustrine sediments, and outwash materials (e.g., silt, sand, gravel, and clay) were deposited by water. Fine-grained material (loess) was deposited by wind, and glacial till was deposited by ice. The last group of parent material is mine spoil, found in areas that have been strip-mined for coal (McNab and Avers 1994). Soil characteristics for each section are described in general terms below, and information on specific soil types is ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Figure 2.—Elevaion zones within Ohio, West Virginia, and Maryland (Danielson and Gesch 2011). available in the U.S. Department of Agriculture, Natural Resources Conservation Service (NRCS) Web soil survey portal. Although anthropogenic activities (e.g., agriculture, mining, urban development) have influenced soil characteristics to some degree in many areas, mountaintop removal mining has been primarily responsible for removal of the soil and bedrock from large areas in southern West Virginia (U.S. Environmental Protection Agency [EPA] 2005, 2011; Wickham et al. 2013). Secion 221F—Western Glaciated Allegheny Plateau Unlike the rest of the assessment area, the geologic history of this section includes major alteration of the land surface caused by the movement of massive glaciers of the Pleistocene Epoch (Ice Age) (Ohio Department of Natural Resources [ODNR] 2005). The land was scoured and depressed under the weight of glaciers, and subsequent melting released the Earth’s crust, deposited boulders and other materials, and carved river beds and other landscape features, such as the rounded hills, ridges, and broad valleys that characterize the area. Glacial features include valley scour, moraines, kames, eskers, and kettle outwash plains. The bedrock consists of shale, siltstone, sandstone, minor conglomerate, and coal (McNab and Avers 1994). The subsurface geology has a generally uniform resistance to erosion, and drainage from smaller streams regularly joins larger streams within a watershed in a regular dendritic pattern. Erosion is caused by two primary geomorphic processes; the first occurs when gravity interacts with moisture on steep slopes to cause slipping of the soil (mass wasting). The second occurs as rivers and streams erode the surrounding earth, transport the soil, and deposit it in a new location (fluvial transport and deposition). Elevation in this section ranges from 50 to 1,500 feet (ODNR 1998). Ridges, flat uplands, hills, and hummocks in this section are dissected by steep valleys covered by thin glacial till and stratified drift. Lower slopes 13 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE and valley floors are covered by sediment and unconsolidated glacial materials. Soils are very deep to bedrock, reaching hundreds of feet in some places (ODNR 2013). The glacial deposits range from coarse-textured to fine-textured, with coarser and better drained soils in the south of this section (ODNR 2013). Lowland surfaces in this section are characterized by gently rolling terrain covered by thin to thick glacial drift with frequent areas of poor drainage and extensive wetlands (ODNR 1998). Secion 221E—Southern Unglaciated Allegheny Plateau This section of the Allegheny Plateau was not covered by glaciers, but was influenced by proximate glaciers as they melted. Fluvial erosion severely dissected the plateau, now characterized by high hills, sharp ridges, and narrow valleys (McNab and Avers 1994). The bedrock is frequently exposed and consists of limestone, siltstone, sandstone, shale, and numerous coal seams. Three major preglacial streams drained the area until many tributaries were blocked by advancing ice sheets, and the accumulation of water formed lakes and deposited sediment in the valleys. This section now has a high density of streams ranging from high gradient, steep headwater streams to low gradient rivers that flow into the Ohio River. Some streams in the preglacial valleys are underlain by relatively shallow silt, sand, or gravel alluvium, and others are filled with deep glacial deposits. Small springs are numerous, but most are ephemeral. Natural streamflow and topography have been greatly modified by oil, gas, and coal extraction activities in this section. Elevation ranges from 490 to 1,400 feet (ODNR 1998). Soils in this section are characterized by a relatively high percentage of clay in the subsoil, associated with remnants of an ancient stream system, where economically important sources of clay and coal are located (ODNR 1998). Clay is found extensively in the lowlands, and red or yellowish-brown silt-loams, 14 silt-clay loam colluvium, and lacustrine silt cover the upland areas (ODNR 1998). Secion M221A—Northern Ridge and Valley This section is characterized by a series of mountain ranges and narrow valleys created by differential erosion of tightly folded, intensely faulted bedrock. The ridges and valleys run parallel from southwest to northeast. From the base of the Blue Ridge Mountains in the east, this section sweeps west across the Great Valley to the Allegheny Mountains and ends at the Allegheny Front, a steep, high ridge marking the eastern boundary of the Allegheny Plateau (McNab and Avers 1994). The bedrock of the ridges consists primarily of resistant sandstone and limestone, and the valleys consist of less resistant shale and siltstone. Erosion and transport of the water-soluble limestone have resulted in sinkholes, caves, and other karst features common in this landscape, and is responsible for the flat topography of the Great Valley (Box 1) (McNab and Avers 1994). Drainage in this section is constricted by the regular folding of bedrock, which forces tributaries to join the main river at right angles in a trellis-shaped pattern. As a result, streams flow in narrow, steep-sided channels (Bruce and Smith 1921). Mass wasting events (landslides), fluvial erosion, and karst solution are common in this section. Elevation ranges from 300 to 4,000 feet. The alluvial soils in this section developed from the weathering of underlying bedrock and the subsequent deposition of sediments laid down in floodplains during stream overflow (Bruce and Smith 1921). Soils are relatively shallow over side slopes, back slopes, and ridges, showing frequent outcropping and escarpments of bedrock (Bruce and Smith 1921, McNab and Avers 1994). The shallowness of the soil is due to the erosion of soil as it forms, the slow formation of soils from highly resistant bedrock, and drier conditions resulting from its position in the rainshadow of the Allegheny Mountains (Bruce and Smith 1921). ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Box 1: Karst Topography Karst landscapes occur where the topography and its disincive features are formed by the dissoluion of soluble rock, especially dolomite and limestone (Fig. 3). The resuling surface features include subterranean drainages, caves, sinkholes, springs, disappearing streams, dry valleys and hollows, natural bridges, arches, and other related features. Sinkholes are karst features that develop as a result of a collapse of surface material into nearby caviies (usually caves). Cold-water springs are characterized by a coninuous low of mineralized groundwater when surface precipitaion percolates through Figure 3.—Potenial karst formaions in the Central Appalachians. Drat map from Weary (2008). Secion M221B—Allegheny Mountains This section is characterized by a series of high, sharp ridges, broad plateaus, low mountains, and narrow valleys created by folded and eroded bedrock (McNab and Avers 1994). The parallel ridges and valleys run southwest to northeast. Bedrock consists of shale, siltstone, limestone, sandstone, and coal. The Greenbrier Limestone generally forms a ring midslope in some mountains, and can be seen as an outcrop in the Canaan Valley in West Virginia. fractures in bedrock including sinkholes, losing streams, caves, and bedrock aquifers. The Northern Ridge and Valley (M221A) and Allegheny Mountains (M221B) secions contain the assessment area’s largest concentraion of soluble rock, therefore the largest karst (Weary 2008). The karst caves are simple or complex subterranean networks that trend northeast along the mountains. Cave and karst systems play an integral role in the area’s biological producivity and provide habitat to rare and endangered species. In West Virginia, 3,754 caves are known and 5 percent of those are reported to support cave-obligate species (Schneider and Culver 2004). Approximately 76 known aquaic and terrestrial species that are dependent on caves are recorded in West Virginia. Several species of state or global viability concern also reside in cave and karst habitats in the assessment area (Byers and Norris 2011). The Indiana bat and Virginia big-eared bat are globally threatened, and populaions of other bat species also depend on these caves. The Carter Cave spider, eastern cave-loving funnel web spider, and Dry Fork Valley cave pseudoscorpion are some of the invertebrate species occurring only in caves, and in some cases, only in a paricular cave in the assessment area (Kovarik 2013). Litle is known about the ecology and life history of many of the cave-dwelling species in the assessment area, making it diicult to determine whether they may be afected by a changing climate. Erosion and transport of the water-soluble limestone has resulted in sinkholes, caves, and other karst features common in this landscape. Drainage is primarily dendritic, but trellis drainage occurs where the topography controls the direction of streamflow. Mass wasting, karst solution, and fluvial erosion are the dominant geomorphic processes. Elevation generally ranges from 1,000 to 4,500 feet, but reaches up to 4,81 feet at Spruce Knob, West Virginia. 15 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE The alluvial soils in this section developed from the weathering of underlying bedrock and the subsequent deposition of sediments laid down in floodplains during stream overflow (Bruce and Smith 1921). Residuum developed mainly from sandstone, shale, and siltstone (Soil Conservation Service [SCS] 1974). Soils that developed in steeply sloping areas are moderately deep, and have a high rock fragment content (35 percent or greater) with moderately well-drained soils in coves (SCS 1974). Gently sloping soils are moderately well-drained and level soils are generally very poorly drained (SCS 1974). Soils also tend to be nutrient-poor and acidic. Soils at higher elevations tend to be shallow with severe soil erosion and lower forest productivity than lower elevation sites. Most of the soils have a severe erosion potential related often to slope but also to other physical properties. Massive soil loss from erosion and wildfires occurred during the logging era (circa 1930s), when thick organic mats were burned to bedrock in places, and sediment filled stream bottoms. Bituminous coal mining has also disturbed large areas of soil, and erosion is associated with stream siltation and acidification. High-elevation soils (above 3,000 feet) receive greater amounts of atmospheric pollutants, especially sulfate (SO4-2) and nitrate (NO3– ), which has led to the loss of important nutrients (e.g., calcium) and the mobilization of others (e.g., aluminum) (Elliott et al. 2013). Secion M221C—Northern Cumberland Mountains This section is characterized by highly dissected uplands and low mountains where less than 20 percent of the area is gently sloping (McNab and Avers 1994). The bedrock consists of shale, coal, sandstone, and limestone. The subsections are named “Eastern Coal Fields” and “Western Coal Fields,” reflecting their realized potential for coal extraction. Drainage is primarily dendritic. Primary geomorphic processes include mass wasting, fluvial erosion, and transport and deposition. Elevation ranges from 00 to 3,900 feet. 1 The soils in this section have formed from residuum on the ridges and mountaintops, colluvium on the slopes, and colluvial and alluvial materials in the valleys. Soils are sandy textured on uplands and loamy on the valley bottoms, where soils are moderately deep to deep and well-drained. Soils can also be shallow and excessively drained in places (NRCS 200). Adequate soil moisture helps to overcome the limitations of these otherwise nutrientpoor soils. hydrology The hydrologic characteristics of the Central Appalachians are influenced, depending on location, by past glaciation, topographic complexity, and proximity to Lake Erie or the Atlantic Ocean. Natural lakes are uncommon in this region; many lakes were created by flooding valleys and building reservoirs (Moore et al. 1997). Many streams have been classified as perennial runoff streams with low baseflow and small variations in year-to-year low flow (Poff 1992). Secion 221F—Western Glaciated Allegheny Plateau This section covers northeastern Ohio, where dominant drainage basins are Lake Erie and the Ohio River. Temporal variations in streamflow primarily follow precipitation and season; snowmelt in early spring increases streamflow, and evapotranspiration during the growing season reduces streamflow. The hydrology of this section is influenced by the remnants of ancient rivers, buried by glacial till. The landscape is characterized by large rivers and floodplains and relatively low topographic relief, with glacial features including kames, kettles, moraines, flat-bottom valleys, bogs, and deranged stream networks. Agriculture and developed lands make up more of the land base in this section than in any other section, placing greater demand on the area’s water resources. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE A shallow slow-moving stream, one of many found throughout the assessment area. Photo by Patricia Butler, Northern Insitute of Applied Climate Science (NIACS) and Michigan Tech, used with permission. Secion 221E—Southern Unglaciated Allegheny Plateau This section covers southeastern Ohio and western West Virginia and contains more than 55,400 miles of streams; 511 lakes, reservoirs, and ponds totaling 27,825 acres; and 12,595 acres of wetlands (Pitchford et al. 2012, West Virginia Department of Environmental Protection [WVDEP] 2013b). Only one lake in West Virginia is natural; the rest were constructed in order to store water. Dominant drainage basins are the Kanawha River and Ohio River basins, which drain westerly to the Gulf of Mexico, and the Potomac River basin, which drains east to the Chesapeake Bay. Streamflow follows the general pattern of growing season and precipitation; intermittent streams are typically dry in late August or early September through early November, when precipitation is low and evapotranspiration is high. Urban and industrial activity is common in valleys along the major rivers. Bituminous coal mining is widespread and has diminished water quality and reduced fish diversity; recent stream quality improvements have occurred in some rivers including the Allegheny, Monongahela, Youghiogheny, and Ohio (Woods et al. 1999). 17 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Secion M221A—Northern Ridge and Valley Dominant drainage basins in this section are the Youghiogheny River draining west to the Ohio River; the Shenandoah River, which joins with the Potomac River; and the Potomac River, which drains east to the Chesapeake Bay. Temporal variations in streamflow primarily follow precipitation and season; snowmelt in early spring increases flow and evapotranspiration during the growing season (June through September) reduces streamflow. Secion M221B—Allegheny Mountains Hydrology varies widely with relief, from flat mountain bogs to steep water gaps. Small ephemeral channels run down ridges to join perennial streams and larger rivers, including the Cheat River, Greenbrier River, and Tygart Valley River. The eastern side of this section is drained by the James River. Steep topography and complex landforms restrict the flow of water so that the swift, actively down-cutting streams run off steep ridges to join the valleys perpendicularly (Woods et al. 1999). Other large rivers such as the Susquehanna River cross ridges, cutting deep gorges in the process (Woods et al. 1999). High-gradient cold-water streams and waterfalls are common in water gaps and on ridge slopes, whereas low-gradient and warmer, meandering streams are common in flatter areas. Because resistant sandstone and shale are not as permeable, surface streams are larger and drainage density is higher than in adjacent limestone areas. Soil erosion is common, and as a result, the stream turbidity can be relatively high and the stream habitat impaired. Bituminous coal mines are common and associated stream siltation and acidification have occurred. Streams do not have much buffering capacity; many reaches, including some not affected by mine drainage, are too acidic to support fish. 18 Secion M221C—Northern Cumberland Mountains Hydrology varies widely with relief, from flat mountain bogs to steep water gaps. Small ephemeral channels run down ridges to join perennial streams and larger rivers, including the Greenbrier River, Guyandotte River, New River, and Tug Fork of the Big Sandy River. This section contains rolling, agricultural lowlands in southeast West Virginia, where limestone bedrock results in karst formation. Stream density is low due to the abundance of saucer-shaped sinkholes. Underground solution channels occur, and subsurface drainage feeds the Greenbrier River. Land Cover The Central Appalachians region is dominated by extensive forests, but also contains other natural ecosystems, rich agricultural lands, urban population centers, and industrial mining lands. Satellite imagery from the National Land Cover Dataset estimates forest cover at 7 percent (Fig. 4) (USGS 2011). The remaining land cover is classified as agricultural land (19 percent), developed land (10 percent), grassland (2 percent), water (1 percent), and wetland (1 percent). Shrublands and barren land (containing no vegetation) make up less than 1 percent of the assessment area. Most of the developed land is located in Ohio, which also has a higher concentration of agricultural lands (crops and hay). A similar concentration of agriculture and development is located in the Great Valley region of Maryland’s panhandle. The relatively small percentage of agricultural land in West Virginia is primarily limited to mountain valleys and gently rolling terrain. Wetlands are scattered throughout the assessment area, occurring over clay soils at the lowest elevations and in geologic depressions at the highest elevations. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Figure 4.—Land cover classes in the assessment area (USGS 2011). DEMoGRAPhiC AND ECoNoMiC CoNDiTioNS There are some important socioeconomic differences among the states within the assessment area. The highest populations are found near Ohio’s lakes and rivers, and in Maryland’s Great Valley. West Virginia remains largely forest, with relatively small urban centers spread throughout the state. Approximately 7.5 million people reside within the assessment area (Headwaters Economics 2011). Seventy-two percent of the population is located in the Ohio portion of the assessment area, with 25 and 3 percent residing in the West Virginia and Maryland portions, respectively. The higher population in Ohio reflects the abundance of lands suitable for agriculture, port access to Lake Erie, and a major shipping industry. The population in the whole assessment area has increased 1 percent since 1970. The population in the Maryland portion of the assessment area has increased the most since 1970 (21 percent). The population in the West Virginia portion has increased by  percent since 1970, and the population in Ohio has decreased by 2 percent. These trends for larger population growth in Maryland are primarily due to the development of Garrett and Washington Counties. The amount of residential acreage increased by 45 percent in Garrett County from 2000 to 2010, largely due to the development of second homes (an increase of 27 percent from 2000 to 2010). The economic well-being of people residing in the assessment area varies across the three states. Unemployment has been highest in the Maryland portion over the past 20 years, but only slightly 19 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE lower in Ohio and West Virginia (Headwaters Economics 2011). In Maryland, growth in employment (57 percent) and personal income (119 percent) over the last 40 years has been greater than in the other two states. Unemployment in the entire assessment area has increased 4 percent since 2007, similar to trends across the United States (Headwaters Economics 2011). Economic Sectors Forest Products industry The forest products industry is measured by grouping census codes from the North American Industry Classification System, and can include logging, forest nurseries, forest products, timber, agriculture and forestry support services, manufacturing of paper and furniture, and other related activities. The forest products industry is important within the assessment area, but the grouping of these codes can vary by state report, and may not be comparable to other states. The forest industry in Maryland is the fifth largest industry in the state; forestry and wood derivatives generated $4.7 billion in 2005 (Maryland Department of Natural Resources [MDNR] 2010). In Ohio, the gross domestic product (GDP) for the manufacturing of wood products and furniture and related products was $2. billion, which represented 0. percent of Ohio’s total GDP in 2007 (ODNR 2010b). The forest products industry in West Virginia contributes approximately $4 billion to the state economy annually (West Virginia Division of Forestry 2010). Forests support jobs and revenue in forest management and logging, sawmills and paper mills, and wood products manufacturing. Within the assessment area, more than 25,000 people are employed in the forest sector, accounting for almost 1 percent of total jobs (Headwaters Economics 2011). The proportion of forestry jobs to total jobs is highest in the Maryland portion (1.9 percent), 20 followed by West Virginia (1.3 percent), and the Ohio portion (0.8 percent). Forestry employment has decreased across the assessment area from 1998 to 2010, which is similar to trends for the United States as a whole (Headwaters Economics 2011). Agriculture The agricultural lands of the assessment area are primarily located where the growing season can last up to 175 days: in the Lake Erie floodplains, Maryland’s Great Valley, and the long, narrow valleys of West Virginia (Woods et al. 199) (Fig. 4). The main agricultural activities are corn, soybean, and beef and dairy livestock farming; Christmas tree plantations; beekeeping; aquaculture; and oilseed and grain farming. Farm employment in the assessment area makes up 1.5 percent of all employment, with a slightly higher percentage of farm employment in West Virginia (2.4 percent) than in other portions of the assessment area. Farm employment has decreased consistently across the assessment area in recent decades, resulting in a net loss of 21 percent of farm jobs from 1970 to 2011 (Headwaters Economics 2011). Muliple land uses in West Virginia. Agriculture and development dominate the lat valleys. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Recreaion The forested lands within the assessment area are a primary destination for recreation opportunities, which are also economically important to the region. The travel and tourism sector generates employment in retail trade, passenger transportation, arts and entertainment, recreation and food, and lodging. In Maryland, increasing demand for forest-based recreation is resulting in increased conflict between motorized and nonmotorized users, and the amount of forest land open to the public is decreasing. Recreation in West Virginia, measured mainly by hunting, fishing, and wildlife viewing receipts, generated $803 million. Tourism in West Virginia generated approximately $4.8 billion, 72 percent of which can be attributed to activities using the state’s forests (West Virginia Division of Forestry 2010). National forests are also a major part of the recreation and tourism industry. Common activities in national forests are viewing natural features, hiking, relaxing, viewing wildlife, driving for pleasure, fishing, motorized trail activity, picnicking, nature study, nature center activities, hunting, gathering forest products, camping, and downhill skiing, among many others (U.S. Forest Service [USFS] 2011b). Travel and tourism provide more than 14 percent of all jobs within the assessment area: 18 percent of jobs within the Maryland portion of the assessment area, followed by West Virginia (1 percent) and Ohio (12 percent) (Headwaters Economics 2011). Travel spending in West Virginia totaled $4.25 billion in 2010, and contributed $988 million in job earnings (Headwaters Economics 2011). Travel-related spending, earnings, and employment have all been increasing in West Virginia since 2000 (Runyan 2011). Total spending on local and nonlocal visits to the two national forests within the assessment area are approximately $50 million per year (USFS 2013). A hiking trail over the Blackwater Falls in the Canaan Valley, West Virginia. Recreaion opportuniies centered around natural resources may also be afected by climate change. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. Mining Mining jobs make up 1.4 percent of all jobs within the assessment area, creating employment in oil and gas extraction, coal mining, metal ore mining, mineral mining, and other related mining work (Headwaters Economics 2011). Coal mining supports the most jobs (27,237 jobs), followed by oil and gas extraction (7,300 jobs), and other related work, such as pipeline construction (3,732 jobs). From 1998 to 2010 mining jobs have decreased in Ohio (-8.2 percent) and Maryland (-19 percent), but increased in West Virginia by 41 percent. Mining is the leading export-oriented industry in West Virginia, generating $.1 billion in GDP in 2009, an increase from its contribution of $5. billion in 2008 (Runyan 2011). Most mining jobs (82 percent) are located in the West Virginia portion of the assessment area (Headwaters Economics 2011). 21 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE ECoSYSTEM CoMPoSiTioN Forest Ecosystems The assessment area is a remarkable landscape of high biodiversity and extensive forests. Many combinations of tree and plant species exist in the variety of habitat conditions that are represented in the area’s 18.9 million acres of forest land (USFS 2013). An ecosystem is defined as a spatially explicit, relatively homogeneous unit of the Earth that includes all interacting organisms and components of the abiotic environment within its boundaries (Society of American Foresters 2014). For the purposes of this assessment, the term “forest ecosystem” refers to a specific classification system and is never used to describe ecosystems in general (Table 2). For example, the Appalachian (hemlock)/ northern hardwood forest is an ecosystem defined by its species assemblage, its spatial distribution, and other distinct characteristics. Common and scientific names of trees and other species mentioned in this assessment are found in Tables 24 through 2 in Appendix 1. Table 2.—Forest classiicaion systems used in this assessmenta Forest ecosystem used in this assessment NatureServe ecological systems represented by the forest ecosystems used in this assessment FiA forest-type groups Common tree species in forest ecosystem Appalachian (hemlock)/ northern hardwood forest Appalachian (hemlock)/northern hardwood forest maple/beech/ birch, aspen/ birch sugar maple, American basswood, American beech, white ash, black cherry, yellow birch, sweet birch, red maple, eastern hemlock, red spruce, tulip tree Dry calcareous forest, woodland, and glade Southern Ridge and Valley/ Cumberland dry calcareous oak/hickory eastern redcedar, chinkapin oak, eastern redbud, eastern hophornbeam, white oak, post oak, shagbark hickory Central Appalachian alkaline glade and woodland North-central Appalachian circumneutral clif and talus Dry oak and oak/pine forest and woodland Allegheny/Cumberland dry oak forest and woodland Central Appalachian dry oak/pine forest Central Appalachian pine/oak rocky woodland oak/hickory, oak/ white oak, black oak, chestnut oak, mockernut hickory, pignut hickory, pine, loblolly/ shortleaf pine scarlet oak, shortleaf pine, pitch pine, Virginia pine, eastern white pine, Table Mountain pine, scrub oak Appalachian shale barrens North-central Appalachian acidic clif and talus (coninued on next page) 22 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Table 2 (coninued). Forest ecosystem used in this assessment Dry/mesic oak forest NatureServe ecological systems represented by the forest ecosystems used in this assessment Northeastern interior dry/mesic oak forest Central and southern Appalachian montane oak forest FiA forest-type groups Common tree species in forest ecosystem oak/hickory, oak/pine, white/ red/jack pine, aspen/birch white oak, black oak, northern red oak, scarlet oak, red maple, pignut hickory, mockernut hickory, shagbark hickory, sugar maple, chestnut oak, sweet birch, American beech, blackgum, tulip tree, white ash oak/gum/ cypress, elm/ ash/cotonwood silver maple, eastern cotonwood, pin oak, red maple, black willow, sycamore, sweetgum, green ash, bur oak, American hornbeam, black walnut, American elm, boxelder, black oak maple/beech/ birch sugar maple, white ash, American basswood, yellow buckeye, tulip tree, red maple, eastern hemlock, cucumbertree, American beech, sweet birch, northern red oak, black cherry, mountain magnolia, black oak maple/beech/ birch sugar maple, American beech, northern red oak, American basswood, eastern hemlock, black cherry, tulip tree, red maple, white ash, eastern hophornbeam elm/ash/ cotonwood sycamore, red maple, silver maple, river birch, boxelder, eastern hemlock, black walnut, pawpaw, American hornbeam, American elm spruce/ir red spruce, yellow birch, eastern hemlock, red maple, sweet birch, cucumbertree, American mountain ash, black cherry, American beech, mountain magnolia, balsam ir, black ash, sugar maple Southern Appalachian oak forest Large stream loodplain and riparian forest South-central interior large loodplain Central Appalachian river loodplain Cumberland riverscour North-central interior lood plain Mixed mesophyic and cove forest South-central interior mesophyic forest Southern and central Appalachian cove forest North-central interior beech/maple forest North-central interior beech/maple forest North-central interior wet latwoods Small stream riparian forest South-central interior small stream and riparian Central Appalachian stream and riparian Cumberland seepage forest Spruce/ir forest Central and southern Appalachian spruce/ir forest Southern Appalachian grass and shrub bald High Allegheny wetland Forest-type groups are used to present broad-scale informaion on forest trends from U.S. Forest Service, Forest Inventory and Analysis (FIA) data. In this assessment, forest ecosystems are used to describe speciic forest communiies and associated environments as commonly grouped by local forest management organizaions. NatureServe (2013) ecological systems were used to describe forest ecosystems, and in many cases, muliple NatureServe systems were combined to describe forest ecosystems within the assessment area. Forest-type groups are classiied diferently from forest ecosystems, and the comparison above is a rough cross-walk between the two systems. a 23 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Forest Classiicaion Systems Used in this Assessment Different organizations describe forests using different classification systems. In this assessment, we describe forests by using two classification systems: (1) USFS Forest Inventory and Analysis (FIA) program (Miles et al. 200) and (2) forest ecosystems, based on NatureServe ecological systems (NatureServe 2013). These classification systems are used for different reasons and convey different types of information. Although there are some general relationships between the two systems, they are organized differently enough that one cannot be substituted for the other. Both types of information are relevant to this assessment; thus, both classification systems are used. FIA data are used to present trends in forest cover, growth, and mortality for forest-type groups, which are defined by tree species composition. Forest ecosystems are also defined by tree species composition, but include associated understory and wildlife species, hydrologic regime, landscape position, and geographic range. FIA forest-type groups are thereby more broadly defined, and can represent several forest ecosystems (Table 2). The FIA program was created by the USFS to characterize forests across the nation. In this assessment, we describe acres, ownership categories, and volume of timber by using “forest-type groups” based on FIA data. FIA classifications describe existing vegetation, and only for vegetated areas dominated by trees (i.e., forests). Forest types are a classification of forest land based upon and named for the dominant tree species. Forest-type groups are a combination of forest types that share closely associated species or site requirements. The FIA system measures tree species composition on a set of systematic plots across the country and uses that information to provide area estimates for each forest-type group. However, it does not make any inferences about what vegetation was historically on the landscape and does not distinguish between 24 naturally occurring and modified conditions. Something that is classified as “forest land” by FIA may have been historically a glade or woodland. Likewise, areas dominated by tree species that are not native to the area would still be assigned to a forest-type group based on dominant species. Throughout this assessment, we also use a classification of “forest ecosystems” as the primary classification system whenever possible because these better describe the forest ecosystems present in the assessment area (Table 2). These forest ecosystems are based upon ecological systems as described by the NatureServe Explorer, a system which is familiar to the Wayne and Monongahela National Forests, state agencies, and other forest management organizations in the assessment area. NatureServe ecological systems describe vegetation as it currently exists on the landscape (NatureServe 2011). An advantage to using the ecological systems is that landforms, soils, and other site features are used whenever possible to help inform the classification. A disadvantage is that the ecological systems have not yet been spatially verified and only rough estimates of abundance are available. In this assessment, we defined our forest ecosystems by combining 24 NatureServe ecological systems into nine forest ecosystems, and then modified the list of dominant species to better reflect the existing vegetation found within the borders of the Central Appalachians assessment area. We used these forest ecosystems to assess vulnerability to climate change (Chapter ). Common and scientific names of trees and other species mentioned in this assessment are found in Tables 24 through 2 in Appendix 1. Forest Ecosystems of the Central Appalachians The following descriptions of forest ecosystems are based on the ecological systems described by the NatureServe Explorer database, which characterizes terrestrial ecosystems at a broad scale across multistate regions (Comer et al. 2003, NatureServe ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE 2013). These systems are classified based on vegetation associations, land cover class, spatial pattern, soil type, and geographic distribution. The assessment area is a region of high biodiversity, and therefore contains more than 40 ecological systems represented within the assessment area boundaries. Of these, we identified the 24 most common forest ecosystems (based on area) and merged systems that were similar or commonly occurred together in order to create nine forest ecosystems that could be assessed by a large panel of experts (Table 2, Appendix 5). Descriptions of each forest ecosystem were further modified to better describe the extent and dominant species of the forest ecosystems as they occur within the boundaries of the assessment area. The resulting forest ecosystems were assessed for their vulnerability to climate change (Chapter ). For original descriptions of NatureServe ecological systems, visit the online database at http://www. natureserve.org/explorer/. Please note that Web addresses are current as of the publication date of this assessment but are subject to change. Appalachian (Hemlock)/Northern Hardwood Forest This forest ecosystem includes only one NatureServe system: the “Appalachian (Hemlock)/Northern Hardwood Forest” system, which extends from southeastern Ohio to the higher-elevation mountains of Maryland and West Virginia. These largely deciduous forests are sometimes mixed with hemlock in the assessment area, distinguishing them from the more montane Southern Appalachian northern hardwood forest. These ecosystems occur on gentle to steep slopes on soils that range from slightly acidic to very acidic with various amounts of nutrients, depending on landscape position and parent material. On colluvial soils, which tend to be on the less acidic end of the spectrum, the canopy is typically dominated by sugar maple, American basswood, American beech, and white ash, with lesser amounts of black cherry, yellow birch, sweet birch, red maple, and tulip tree. Sites on the more acidic end of the spectrum are usually dominated by combinations of yellow birch, American beech, black cherry, red maple, and eastern hemlock, although most sites will not have all five species present as canopy dominants. Minor components on the more acidic sites include sweet birch, red spruce, and tulip tree. On both colluvial and slightly less acidic soils, sweet birch and tulip tree may become dominant in response to heavy disturbance. Wind-driven gap disturbances are the most influential natural disturbance. Historically, this forest ecosystem was probably subject to only extremely rare natural fires. Logging and subsequent slash fires around the turn of the 20th century probably promoted deciduous species at the expense of hemlock and red spruce. Consequently, many examples of this forest ecosystem on more acidic sites are likely successional to the spruce/fir forest ecosystem. The eastern hemlock component is currently being decimated in large parts of its range by the hemlock woolly adelgid, which will likely result in replacement of hemlocks by other canopy trees (Hessl and Pederson 2013). Likewise, the American beech component is being decimated by beech bark disease, which typically results in conversion to broken-canopied stands with a dense understory of beech sucker sprouts. An Appalachian (hemlock)/northern hardwood forest. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 25 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Dry Calcareous Forest, Woodland, and Glade This forest ecosystem is mostly formed by small patches of three NatureServe (2013) systems that occur on thinner circumneutral and calcareous soils over limestone or dolostone. The “Central Appalachian Alkaline Glade and Woodland” system occurs on ridges, summits, and upper to lower slopes and usually grades into closed-canopy forests at low to moderate elevations. Common tree species include eastern redcedar, chinkapin oak, eastern redbud, and eastern hophornbeam. Prairie grasses and forbs dominate the herbaceous layer, with a number of rare forbs. Most existing open patches appear to be maintained by drought and landslides. Fire frequency and intensity influence the relative ratio of deciduous to evergreen trees, with eastern redcedar increasing in the absence of fire (Smith and Johnson 2004). Whether the woodland portion occupied larger areas under a historic regime of frequent fire is debatable. Soils are excessively welldrained, with low water holding capacity. The open canopies and relatively high pH soils make these communities more susceptible to invasive species. On deeper soils in extreme southeast West Virginia, the “Southern Ridge and Valley/Cumberland Dry Calcareous Forest” system may be dominated by white oak and shagbark hickory, and sometimes contains eastern redcedar as a significant component. The “North-Central Appalachian Circumneutral Cliff and Talus” system occurs in small patches on vertical or near-vertical cliffs and steep talus slopes at low to moderate elevations with alkaline soils. Lichens are generally the most abundant vegetation, although stunted northern white cedar may occur on north-facing cliffs. The historic natural disturbance regime included frequent low-intensity fires, but contemporary disturbances include exposure and landslide events. Dry Oak and Oak/Pine Forest and Woodland This forest ecosystem includes major patchforming forests and woodlands. NatureServe (2013) describes this forest as the “Allegheny/Cumberland 2 Dry Oak Forest and Woodland” on predominantly acidic substrates on southwest-facing slopes in the Allegheny Plateau, where it is dominated by white oak, black oak, chestnut oak, mockernut hickory, pignut hickory, and scarlet oak with small inclusions of shortleaf pine and Virginia pine. The “Central Appalachian Dry Oak/Pine Forest” (NatureServe 2013) occurs in the rainshadow areas of the Ridge and Valley section at low to high elevations, and is dominated by a variable mixture of dry-site oak and pine species including chestnut oak, white oak, scarlet oak, pitch pine, Virginia pine, Table Mountain pine, and eastern white pine. Ericaceous shrubs are common in the understory of both systems. On wooded hilltops, outcrops, cliff faces, and rocky slopes, NatureServe (2013) classifies this forest as “Central Appalachian Pine/Oak Rocky Woodland” and “North-Central Appalachian Acidic Cliff and Talus” systems. These smaller patches are dominated by lichens and stunted trees and may form a woodland with pitch, Virginia, and Table Mountain pines mixed with xerophytic oak species and sprouts of American chestnut. Another patch system, “Appalachian Shale Barrens” (NatureServe 2013), occurs where exposed shale creates extreme growing conditions. Many of these patches occur as open land or as woodland dominated by stunted chestnut oak, Virginia pine, pignut hickory, and scrub oak. As many as 15 endemic herbaceous species may occur on these shale systems. Soils are generally xeric and sandy, and have low water holding capacity. Fire was historically frequent in this forest ecosystem, but contemporary fire suppression has led to shifts in species composition and stand structure. Dry/Mesic Oak Forest This forest ecosystem includes two matrix-forming oak-dominated systems that are only weakly differentiated and occupies more area than any other forest ecosystem in the assessment area. NatureServe (2013) describes this forest as “Northeastern Interior Dry/Mesic Oak Forest” in the north, “Southern Appalachian Oak Forest” in the south, and “Central ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE and Southern Appalachian Montane Oak Forest” in the Allegheny Mountains. The dry/mesic oak forests are drier than the mixed mesophytic and cove forest and more mesic compared to the dry oak and oak/pine forest and woodland. This system is often stunted and wind-flagged on exposed southwest slopes and ridge crests. Common species include white oak, northern red oak, black oak, and scarlet oak. Associated canopy trees also include red maple, pignut hickory, mockernut hickory, shagbark hickory, sugar maple, chestnut oak, tulip tree, sweet birch, white ash, American beech, and blackgum. American chestnut and eastern white pine were historically dominant or codominant species in some areas. Fire is an important driver in this forest ecosystem, but contemporary fire suppression has favored maple species over oaks. Wind and ice storms continue to be important disturbances. Large Stream Floodplain and Riparian Forest This forest ecosystem is found across the assessment area as a complex of wetland and upland vegetation associated with medium to large rivers or streams where topography and alluvial processes have resulted in a well-developed floodplain. NatureServe (2013) describes these forests as “Central Appalachian River Floodplain,” “Cumberland Riverscour,” and “North-Central Interior Floodplain.” There is typically a gradient from moist or periodically dry, somewhat nutrient-enriched conditions upslope to moist and highly enriched conditions downslope. Most areas are inundated with seasonal flooding, most commonly in the spring; microtopography determines how long the various habitats are inundated. Seasonal flooding and floodscouring contribute to sediment deposition and can be abrasive forces along the riverbanks. Some Yellow birch and hemlock, common riparian species in the Allegheny Mountains. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 27 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE areas are also prone to severe drought periods that may stress or kill vegetation. A variety of alluvial and loess soil types, in combination with various flooding regimes, create a diversity of vegetation communities such as floodplain forests, herbaceous sloughs, shrub wetlands, riverside prairies, and woodlands. The wettest areas are dominated by silver maple, eastern cottonwood, pin oak, red maple, and black willow. Better-drained areas are dominated by sycamore, sweetgum, green ash, bur oak, American hornbeam, black walnut, American elm, boxelder, and black oak. Some common shrubs are hazel alder, common buttonbush, silky dogwood, coastal plain willow, pawpaw, spicebush, and eastern redcedar. Anthropogenic land conversion and invasive plant species are major stressors in this forest ecosystem. Mixed Mesophyic and Cove Forest This forest ecosystem is located entirely south of the glacial boundary, and it is predominantly found in West Virginia within the assessment area. The Allegheny Front separates two similar systems: NatureServe (2013) describes this system as “SouthCentral Interior Mesophytic Forest” in the west, and “Southern and Central Appalachian Cove Forest” in the east. This forest ecosystem consists of mesophytic hardwood or hemlock-hardwood forests in sheltered topographic positions, often on concave slopes or in areas with high precipitation. Common species include sugar maple, white ash, American basswood, yellow buckeye, tulip tree, red maple, eastern hemlock, American beech, cucumbertree, sweet birch, northern red oak, black cherry, and mountain magnolia. Black oak and black walnut can also occur as minor canopy species. Soils are predominantly colluvial, and range from slightly basic to very acidic, with various amounts of nutrients. Rich coves collect moisture and nutrients from higher positions, and support higher 28 diversity and density in the herbaceous layer and tree layer. Acidic coves often have a dense shrub layer dominated by great laurel and mountain laurel. This system is naturally dominated by unevenaged forests, with gap-phase regeneration, although current conditions resemble more even-aged second-growth forests. Occasional extreme wind or ice events may disturb larger patches. Natural fires are probably extremely rare and have occurred only in years that were extremely dry. Most of the component species are among the least firetolerant in the region. Trees may grow very large in undisturbed areas, but repeated harvesting can result in smaller age-class distributions and favor tulip tree and red maple. North-Central Interior Beech/Maple Forest This forest ecosystem includes the NatureServe (2013) “North-Central Interior Beech/Maple Forest” system on gently rolling uplands to moderate slopes, and the “North-Central Interior Wet Flatwoods” on poorly drained uplands or in clay-lined glacial depressions. This forest ecosystem is primarily found in the glaciated portion of Ohio, where various microtopography and moisture regimes create mixed communities of upland and lowland species. These forests can be composed of deciduous or mixed evergreen-deciduous species including sugar maple, American beech, northern red oak, American basswood, eastern hemlock, black cherry, tulip tree, red maple, white ash, and eastern hophornbeam. On upland sites, soils are loamy over glacial till, limestone, or calcareous shales, and have adequate or abundant levels of nutrients. In wetter locations, soils typically have an impermeable clay layer resulting in ponding and complete saturation during spring and possible drought during summer. The disturbance interval is long, with wind as the primary disturbance, and this forest ecosystem is generally intolerant of fire. ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Small Stream Riparian Forest This forest ecosystem is a matrix of uplands and wetlands found along creeks, small streams, and medium rivers (e.g., Shaver’s Fork) with low to moderately high gradients and oxbows. NatureServe (2013) systems include “South-Central Interior Small Stream and Riparian,” “Central Appalachian Stream and Riparian,” and “Cumberland Seepage Forest.” Flooding and scouring both influence this system, but the nature of the landscape (i.e., steeper side slopes and higher gradients) prevents the kind of floodplain development found along larger rivers. Soils are colluvial and alluvial deposits with moderate inherent fertility, ranging from moist to periodically dry (i.e., poorly to excessively well-drained). The vegetation is a mosaic of forests, woodlands, shrublands, and herbaceous communities. Typical tree species may include sycamore, red maple, silver maple, river birch, boxelder, eastern hemlock, black walnut, pawpaw, American hornbeam, and American elm, as well as many of the tree species that occur in adjacent upland forests. The eastern hemlock component is threatened by the hemlock woolly adelgid, which will likely result in its replacement by other canopy trees. Some characteristic shrubs may include bushy St. Johnswort, coastal plain willow, and hazel alder. Spruce/Fir Forest This forest ecosystem consists of forests, woody wetlands, shrublands, and grasslands on a variety of landforms in the highest elevation zone of the Allegheny Mountains, ranging from 2,400 to 4,00 feet. NatureServe (2013) describes these forests as “Central and Southern Appalachian Spruce/Fir Forest,” “Southern Appalachian Grass and Shrub Bald,” and “High Allegheny Wetland.” Elevation and topography make the climate cool and wet, with heavy moisture input from fog as well as high amounts of rain and snow. Soils are moist year-round, usually acidic, and often very rocky, originating from weathered parent material or from organic deposits over boulders. The forest canopy is typically dominated or codominated by A red spruce and mixed hardwood forest in West Virginia. Photo by David Ede (reired), Monongahela Naional Forest. red spruce, with associates including yellow birch, red maple, and eastern hemlock. In some places, sweet birch, cucumbertree, American mountain ash, black cherry, American beech, sugar maple, and mountain magnolia may also appear. The eastern hemlock component is currently being decimated in large parts of its range by the hemlock woolly adelgid, which will likely result in replacement of hemlocks by other canopy trees. Likewise, the American beech component is being decimated by beech bark disease, which typically results in a dense understory of beech sucker sprouts. Balsam fir and black ash can also dominate in wet areas on limestone or calcareous shale. On upland sites, the shrub layer can range from sparse to dense and may include great laurel and southern mountain cranberry. Around the edges of some wetlands, the shrub layer may be dense and may contain a variety of species, including wild-raisin, velvetleaf huckleberry, speckled alder, bushy St. Johnswort, common winterberry, and black chokeberry. The herbaceous layer is generally sparse in upland areas and dense in wetlands. Fine-scale disturbances (e.g., debris avalanches, wind, and ice) are generally the most influential in this forest ecosystem. Red spruce and eastern hemlock are both expanding into portions of their historic niches, recovering from large anthropogenic disturbances at the beginning of the last century. 29 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Forest Composiion and Abundance Analysis of satellite imagery from the National Land Cover Database estimates forest coverage at 7 percent of the land base. The FIA program, using a network of permanent field plots, estimates that  percent of land is forested (Table 3) (USFS 2013). Northern Ohio (221F) is the least forested section (31 percent), and southern West Virginia (M221C) is the most heavily forested section (89 percent). Timberland is forest land that is currently producing or capable of producing more than 20 cubic feet of wood per acre per year. Approximately 97 percent of the forest land in the assessment area is classified as timberland (USFS 2013). Based on FIA data, the oak-hickory forest-type group is the most common in the assessment area, covering 70 percent of the total forested area (Table 4). The other common forest-type groups across the assessment area are the maple/beech/birch group (19 percent), elm/ash/cottonwood group Table 3.—Acreage (total and forest land) for each ecological secion within the assessment area, as determined by FIA (USFS 2013) Ecological secion Total land (acres) Forest land Forest land (acres) (%) 221E 221F M221A M221B M221C 13,485,413 4,961,053 2,512,194 4,374,448 3,504,533 9,056,910 1,561,695 1,778,653 3,366,526 3,130,745 67 31 71 77 89 Assessment area 28,837,641 18,894,529 66 (4 percent), oak/pine group (2 percent), and loblolly/ shortleaf pine group (1 percent). The remaining forest-type groups each equal less than 1 percent of the forest land. There are also more than 8,000 acres of nonnative blue spruce plantation that was classified as the fir/spruce/mountain hemlock foresttype group. Differences among forest types can influence the amount of carbon stored aboveground and belowground (Box 2). Table 4.—Forest land by FIA forest-type group (USFS 2013) FIA forest-type group Total assessment area (acres) Total assessment area (%) Oak/hickory Maple/beech/birch Elm/ash/cotonwood Oak/pine Loblolly/shortleaf pine White/red/jack pine Other hardwoods Nonstocked Aspen/birch Exoic hardwoods Spruce/ir group Exoic sotwoods group Other eastern sotwoods Oak/gum/cypress Fir/spruce/mountain hemlock 13,311,652 3,601,227 667,362 430,319 241,679 162,162 141,683 98,697 85,235 43,649 40,368 22,031 20,945 19,376 8,144 70.5 19.1 3.5 2.3 1.3 0.9 0.7 0.5 0.5 0.2 0.2 0.1 0.1 0.1 0.0 Total 18,894,529 30 100 Ecological secion within the assessment area (acres) 221E 221F M221A M221B M221C 6,602,077 1,461,080 392,774 232,885 135,343 62,503 23,396 62,566 54,007 30,279 – – – – – 828,364 424,142 196,052 – – 12,050 12,611 21,662 21,024 6,292 – 18,920 – 12,434 8,144 1,328,991 1,926,512 2,625,709 149,336 1,154,314 412,356 33,296 5,470 39,770 111,333 54,373 31,729 83,496 16,179 6,660 27,842 57,692 2,075 25,320 74,987 5,370 6,397 8,072 – – 10,204 – – – 7,078 – 40,368 – – 3,111 – 12,642 8,302 – – 6,942 – – – – 9,056,910 1,561,695 1,778,653 3,366,526 3,130,747 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Box 2: Forest Carbon in the Assessment Area Forests play a valuable role as carbon sinks. The accumulated terrestrial carbon pool within forest soils, belowground biomass, dead wood, aboveground live biomass, and liter represents an enormous store of carbon (Birdsey et al. 2006). Terrestrial carbon stocks in the region have generally been increasing for the past few decades, and the potenial for managing forests to maximize and maintain this carbon is gaining atenion (Malmsheimer et al. 2011). Carbon sequestraion and storage in forest ecosystems depend on the health and funcion of those ecosystems in addiion to human management, episodic disturbances, and forest stressors. Forest lands within the assessment area are esimated to hold approximately 1.3 billion metric tons of carbon, or roughly 69.1 metric tons per acre (USFS 2013). Depending on the forest-type group, carbon density ranges from 42.7 metric tons per acre (other eastern sotwoods group) to 98.6 metric tons per acre (oak/gum/cypress group) (Fig. 5). The spruce/ir and maple/beech/birch groups store greater amounts of carbon per acre than the oak/hickory and oak/pine groups. However, because the vast majority of forest land is classiied as oak/hickory and maple/beech/birch, most of the total carbon in the assessment area is found in these two types (67 percent in oak/hickory and 22 percent in maple/beech/birch). Carbon density also varies by ownership. The highest density of carbon is on federal lands administered by the USFS, the Naional Park Service, the U.S. Fish and Wildlife Service (USFWS), and the Department of Defense, ranging from 76.5 to 87.7 metric tons per acre. Private lands store only 67.7 metric tons per acre. However, most of the forest land in the assessment is private land. Therefore, 1 billion metric tons of carbon is stored on private land, versus 209 million metric tons stored on public lands. 100 C arb o n d en sity (m etric to n s p er acre) 90 80 70 60 50 40 30 20 S o il o rg anic Litter D ead w o o d Liv e b elo w g ro und Liv e ab o v eg ro und 10 0 Figure 5.—Forest carbon density by forest-type group. Forest-type groups are arranged from let to right by area of forest land (USFS 2013). 31 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE DRiVERS oF ChANGE iN FoREST ECoSYSTEMS The forest ecosystems of the assessment area have undergone significant changes over the past several thousand years. These changes were largely driven by periodic climate change and anthropogenic pressures on the natural resource base, which in turn have had major implications for fire occurrence and behavior, invasive species establishment, soil stability and structure, hydrology, and other drivers of species composition and structure. Past Ecosystem Change Paleoecological records from pollen and macrofossils have been collected from lakes and bogs throughout the eastern United States as a means to determine long-term vegetation dynamics (Davis 1983). During the last glacial maximum 18,000 to 20,000 years before present (YBP), only a portion of the assessment area (Section 221F in Ohio) was covered in ice. In the rest of the assessment area, a belt of tundra extended from Pennsylvania southward along the mountains. Forest species formed novel assemblages of conifers such as spruce and pine, and deciduous species were conspicuously absent (Davis 1983). After the last glaciers began to retreat from the northern latitudes, at the start of the Holocene epoch approximately 11,700 YBP, tree species started to respond to the warming climate and the melting glaciers. Many tree species were able to migrate northward at a rate of 700 to 1,000 feet per year, but expansion over the Appalachian Mountains was generally slower (around 300 feet per year) (Davis 1983). Spruce and fir moved northward at different rates, depending on the suitability of climate, seed dispersal, and establishment success (Davis 1983). The large-scale replacement of eastern white pine and other northern species by deciduous or mixed deciduous forest occurred between 10,000 and 12,000 YBP (Jacobson et al. 1987). Elms and maples arrived from the west around 10,000 to 12,000 YBP. Oaks arrived 10,000 32 YBP and dominated the landscape until 500 YBP. Some species arrived relatively recently in the assessment area, largely due to the migration barrier presented by the Appalachian Mountains. Hickories arrived 10,000 YBP in Ohio, but expanded slowly over the Appalachian Mountains, reaching Maryland only 5,000 YBP. Chestnut arrived in Tennessee 10,000 YBP, and reached Maine after a slow northeast migration only 2,000 YBP. Eastern white pine and eastern hemlock expanded from ancient forest refugia near the Atlantic coast and arrived 10,000 YBP (Davis 1983). Repeated periods of warming and cooling over the last 15,000 years have resulted in multiple waves of species expanding from the south and from climatic refuges along the Atlantic coast (Shuman et al. 2002). These waves of species migrations resulted in very different species assemblages from those typical today, partially because not all species were able to migrate at the same rate. The last major shift in climate occurred approximately 3,000 YBP. This wetter and cooler climate is similar to our present climate, which favors tree growth and reforestation. At the same time, the effects of Native Americans on the vegetation of the region became evident. Charcoal scars throughout the region have confirmed a link between oak dominance and fire (Abrams 1992, Nowacki and Abrams 2008). Native Americans are thought to be responsible for the numerous low-intensity fires that promoted oak regeneration (Abrams 1992). Native American cultures centered on maize agriculture were in place by 1,000 YBP. The development of smallscale agriculture and other activities also resulted in extensive trail and trade networks and subsistencebased manipulation of the vegetation. By the early 17th century, the use of fire by Native Americans began to diminish as native populations crashed from disease and as European settlers laid claim to land. Witness trees, as recorded in surveyors’ notes, are often the only indication of what composed ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE presettlement forests and their disturbance regimes (Black and Abrams 2001). When grouped by fire relations, witness trees can be converted to show spatial differences in presettlement fire regimes (Thomas-Van Gundy and Nowacki 2013). Witness tree data show that white oak was dominant over large areas of the Central Appalachians (Table 5) (Abrams 2003). Exceptions occurred in Ohio’s glaciated plateau, where white oak was codominant in maple-beech forests, and in the many landforms of the Appalachian Mountains. Analysis of witness trees within the Monongahela National Forest correlates white oak with low elevation and high moisture, whereas higher elevations supported sugar maple, American beech, birch, red spruce, eastern hemlock, and black cherry, among others (ThomasVan Gundy and Strager 2012). Industrialization and settlement during the 18th and 19th centuries also created heavy demands on forests within the Central Appalachians region. The logging boom of 1880 to 1930 is often considered the most important driver of forest ecosystems in the assessment area, although the forests of Ohio had already declined from 95 percent of land cover to 40 percent by 1880, largely due to agriculture (Birch and Wharton 1982, Widmann et al. 2007). Large-scale clearcutting was conducted for the purposes of wood harvesting and agricultural land clearing. The effects of repeated logging that removed most old-growth forests, and the subsequent wildfires, are still being observed today. Secondary forests are largely even-aged with poor structure and reduced species diversity. Other impacts include the loss of soil that will take thousands of years to replace, degraded stream channels, and old railroad grades and logging roads that impair watershed hydrology and create edge effects. In the early 1900s, frequent and intense fires favored oaks, hickories, and yellow pines at the expense of hemlock, red spruce, white pine, and mesophytic hardwoods. Table 5.—Witness tree observaions from various locaions in the Central Appalachians Region, locaion Presetlement forest composiion Reference Southwestern Pennsylvania White oak (40%), black oak (9%), hickory (9%), dogwood (8%) (Abrams and Downs 1990) Eastern West Virginia Ridges White oak (35%), chestnut (15%), chestnut oak (13%), black oak (12%) (Abrams and McCay 1996) Eastern West Virginia Valleys White oak (23%), maple (22%), pine (15%), basswood (10%) (Abrams and McCay 1996) Southern West Virginia White oak (24%), chestnut (12%), hickory (9%), (Abrams et al. 1995) chestnut oak (6%) Monongahela Naional Forest White oak (19%), sugar maple (10%), American beech (8%) (Thomas Van-Gundy and Strager 2012) Southeastern Ohio White oak (40%), hickory (14%), black oak (12%), American beech (8%) (Dyer 2001) Northeastern Ohio Fine ill American beech (36%), sugar maple (17%), white oak (14%) (Whitney 1994) Northeastern Ohio Coarse ill White oak (37%), hickory (13%), black oak (6%) (Whitney 1994) 33 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Primary Agents of Change Agents of change within the assessment area include both natural and anthropogenic pressures. Fire suppression, wind events, severe weather, pests and diseases, invasive species, large-scale surface mining, acid deposition, fragmentation, and land use change are the primary agents of change in the Central Appalachians region. Each of the forest ecosystems addressed in this assessment faces a particular suite of threats and stressors (Table ). We define threats and stressors as agents that tend to disrupt the natural functioning of forest ecosystems or impair their health and productivity. This information is collected from published literature as well as local forest managers. The impacts of particular threats and stressors are very dependent on local conditions and are not consistent across an area as large and diverse as the Central Appalachians. These particular threats should be considered in addition to landscape-level threats such as acid deposition, forest fragmentation, the legacy of past management practices, and altered disturbance regimes. It is often difficult to examine the effects of just one of these landscape-level threats in isolation, because they have all interacted across the assessment area over the past century. Fragmentation caused by mining, agricultural and urban development, forest management, and other factors has tended to reduce the ratio of interior to edge conditions in forests (Drohan et al. 2012a, Irwin and Bockstael 2007). The disruption of natural disturbance regimes has included fire suppression in upland systems as well as hydrologic disruption in riparian and lowland forests. Natural regeneration and succession of forest ecosystems is strongly tied to disturbance regimes, so in many cases alteration of disturbance regimes has resulted in regeneration failure for those disturbance-adapted species and reduced landscape diversity (Abrams and Nowacki 1992, Nowacki and Abrams 2008, Patterson 200). Conversely, other species may benefit from the altered disturbance regime, particularly firesensitive, shade-tolerant trees. Table 6.—Major disturbances and threats to forest ecosystems in the Central Appalachians Forest Ecosystem References All forest ecosystems (Central Appalachians) Atmospheric deposiion of nitrates, sulfates, ozone, and other anthropogenic emissions negaively afects forest producivity and resilience. (Poter et al. 2010) Deer herbivory is considered a keystone driver through impacts on plant regeneraion, structure, and species diversity, especially where deer density is high. (Collins and Carson 2003, MDNR 2010, ODNR 2010b) Drought can lead to increased ire hazard, decreased plant growth, regeneraion failure, and increased suscepibility to insects and diseases. (ODNR 2010b) Energy development for wind energy and shale-gas installaions alter ecosystem structure through vegetaion clearing, soil disturbance, increased erosion potenial, fragmentaion, and direct impacts on forest wildlife species. (Drohan et al. 2012b, Naional Research Council 2007) Fragmentaion from industrial and urban development has resulted in dispersal barriers that impede migraion of species and exchange of geneic material, reduced forest patch size, and increased forest edge. (Irwin and Bockstael 2007, Poter et al. 2010, Riiters 2011) Geographic dispersal barriers slow the dispersal and migraion of species in muliple direcions across the Appalachian Mountains. (Davis 1983) insect pests and diseases increase the risk of individual tree mortality and species exincion or exirpaion. (DeSanis et al. 2013, Lovet et al. 2006, MDNR 2010, ODNR 2010b, Poter et al. 2010) (coninued on next page) 34 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Table 6 (coninued). Forest Ecosystem References Appalachian (Hemlock)/Northern Hardwood Forest Acid deposiion negaively afects forest producivity and resilience. (ODNR 2010b, Poter et al. 2010) Deer herbivory results in reduced growth and mortality of seedlings and saplings of target browse species. (Collins and Carson 2003, MDNR 2010) insect pests and diseases such as emerald ash borer, hemlock woolly adelgid, hypoxylon canker, and beech bark disease can cause reduced growth and mortality of target species. (Anderson 1995, Burns and Honkala 1990, DeSanis et al. 2013, Hessl and Pederson 2013, MDNR 2010) Invasive plants such as garlic mustard, ailanthus, Japanese siltgrass, basket grass, and paper mulberry reduce natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) Dry Calcareous Forest, Woodland, and Glade insect pests and diseases such as red oak borer, gypsy moth, sudden oak death, oak decline, and armillaria root disease can cause reduced growth and mortality of target species. (MDNR 2010, ODNR 2010b) Invasive plants such as ailanthus, Asiaic bitersweet, garlic mustard, mulilora rose, Japanese honeysuckle, bush honeysuckle, autumn olive, spoted knapweed, viper’s bugloss, Japanese siltgrass, and Canada bluegrass can reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Hutchinson and Vankat 1998, Kurtz 2013) Suppression of natural ire regimes has contributed to woody encroachment of eastern redcedar and mountain laurel; overabundance of these shrubs can reduce diversity and afect species regeneraion. (Abrams 1992, Nowacki and Abrams 2008, Smith and Johnson 2004) Dry Oak and Oak/Pine Forest and Woodland insect pests and diseases such as gypsy moth, sirex woodwasp, southern pine beetle, oak decline, and armillaria root disease can cause reduced growth and mortality of target species. (MDNR 2010, ODNR 2010b) Invasive plants such as ailanthus, Japanese siltgrass, mulilora rose, Japanese honeysuckle, bush honeysuckle, autumn olive, Japanese barberry, sericea lespideza, yellow sweetclover, and crown vetch reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) Past management aciviies which created microsite condiions conducive to pine regeneraion are diicult to reconstruct without intense ire, resuling in a gradual conversion from pine to oak species. (Vose et al. 1993) Suppression of natural ire regimes has reduced structural and species diversity, allowed mesic hardwood encroachment on many sites, and limited suitable condiions for natural regeneraion. (Abrams 2003, Nowacki and Abrams 2008, Paterson 2006, Sharitz et al. 1992) (coninued on next page) 35 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Table 6 (coninued). Forest Ecosystem References Dry/Mesic Oak Forest insect pests and diseases such as ambrosia beetle, red oak borer, gypsy moth, oak decline, (Lovet et al. 2006, ODNR armillaria root disease, hypoxylon canker, and phytophtora root rot can cause reduced 2010b, Rafa et al. 2008) growth and mortality of target species. Invasive plants such as ailanthus, Japanese siltgrass, and garlic mustard reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013, MDNR 2010) Suppression of natural ire regimes has reduced structural and species diversity, allowed mesic hardwood encroachment on many sites, and limited suitable condiions for natural regeneraion. (Abrams 2003, Nowacki and Abrams 2008, Paterson 2006, Sharitz et al. 1992) Large Stream Floodplain and Riparian Forest Energy development for wind energy and shale-gas installaions alter ecosystem structure through vegetaion clearing, soil disturbance, increased erosion potenial, polluion, fragmentaion, mine land abandonment, and direct impacts on forest wildlife species. (Drohan et al. 2012b, Naional Research Council 2007) Erosion from improperly designed or poorly maintained roads, trails, or log landings can (Pennsylvania Department increase the amount of siltaion and sedimentaion transported and deposited by streams. of Environmental Protecion 2012) Industrial and urban development has resulted in hydrologic infrastructure that afects the lood regime, such as impoundments, channelizaion, dams and reservoirs, and drainage and clearing for agriculture. (Irwin and Bockstael 2007, Poter et al. 2010, Riiters 2011) insect pests and diseases such as emerald ash borer, thousand cankers disease, and elm yellows can cause reduced growth and mortality of target species. (DeSanis et al. 2013, Graton 2013, Kurtz 2013) Invasive plants are transported by water and establish more rapidly here than in other systems. Species such as Japanese siltgrass, Japanese hops, and bush honeysuckle can reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) Mixed Mesophyic and Cove Forest Forest arson and debris burning are the major causes of wildire. (MDNR 2010, ODNR 2010b) insect pests and diseases such as emerald ash borer, hemlock woolly adelgid, and beech bark disease can cause reduced growth and mortality of target species. (Hessl and Pederson 2013, ODNR 2010b) Invasive plants such as Japanese siltgrass, garlic mustard, ailanthus, and bush honeysuckle can reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) Mountaintop removal mining and valley ill changes topography, soil water capacity, and runof; and buries headwater streams where mining waste is dumped. (U.S. EPA 2005, 2009) (coninued on next page) 3 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Table 6 (coninued). Forest Ecosystem References North-Central Interior Beech/Maple Forest Energy development for wind energy and shale-gas installaions alter ecosystem structure through vegetaion clearing, soil disturbance, increased erosion potenial, fragmentaion, and direct impacts on forest wildlife species. (Drohan et al. 2012b, Naional Research Council 2007) Fragmentaion and urban development has resulted in dispersal barriers that impede migraion of species and exchange of geneic material, reduced forest patch size, and increased forest edge. (Irwin and Bockstael 2007, Poter et al. 2010, Riiters 2011) Invasive plants such as princess tree, silk tree, garlic mustard, creeping charlie, Japanese siltgrass, ailanthus, and glossy buckthorn reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) insect pests and diseases such as hemlock woolly adelgid, gypsy moth, emerald ash borer, sirex woodwasp, anthracnose disease, sudden oak death, and beech bark disease can cause reduced growth and mortality of target species. (DeSanis et al. 2013, Hessl and Pederson 2013, ODNR 2010b) Small Stream Riparian Forest Energy development for wind energy and shale-gas installaions alter ecosystem structure through vegetaion clearing, soil disturbance, increased erosion potenial, polluion, fragmentaion, mine land abandonment, and direct impacts on forest wildlife species. (Drohan et al. 2012b, Naional Research Council 2007) Erosion from improperly designed or poorly maintained roads, trails, or log landings can (Pennsylvania Department increase the amount of siltaion and sedimentaion transported and deposited by streams. of Environmental Protecion 2012) Industrial and urban development has resulted in hydrologic infrastructure that afects the lood regime, such as impoundments, channelizaion, dams and reservoirs, and drainage and clearing for agriculture. (Irwin and Bockstael 2007, Poter et al. 2010, Riiters 2011) Invasive plants are transported by water and establish more rapidly here than in other systems. Species such as Japanese siltgrass, Japanese hops, and bush honeysuckle can reduce suitable condiions for natural regeneraion, facilitate other exoic species, and alter understory plant communiies. (Graton 2013, Kurtz 2013) insect pests and diseases such as emerald ash borer, hemlock woolly adelgid, thousand cankers disease, and elm yellows cause reduced growth or mortality of target species. (DeSanis et al. 2013, Graton 2013, Kurtz 2013) Spruce/Fir Forest Acid deposiion at high-elevaion sites adversely afects the growth and physiology of red (Friedland et al. 1984, spruce. Acid deposiion has also been linked to increased predisposiion to frost damage in McLaughlin et al. 1990, red spruce. Schuler and Collins 2002) Anthropogenic impacts from surface mining and wind energy development, roads, (Schuler and Collins 2002) recreaion, and residenial development have resulted in fragmentaion, altered hydrology, and forest conversion. Deer browse results in reduced growth and mortality of seedlings and saplings of target browse species (e.g., eastern hemlock). (Michael 1992, Schuler and Collins 2002) Frost damage is a major cause of foliar loss in red spruce. (Friedland et al. 1984) insect pests and diseases such as hemlock and balsam woolly adelgids, emerald ash borer, (DeSanis et al. 2013, Hessl and beech bark disease can cause reduced growth and mortality of target species. and Pederson 2013, Schuler and Collins 2002) 37 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Fragmentaion and Land-use Change Residential and urban development has led to the fragmentation of forests across the assessment area, resulting in a patchwork of public and private parcels of natural, agricultural, and developed lands. As mentioned earlier, 40 percent of the assessment area is now agricultural or developed land (USFS 2013). Northern Ohio and western Maryland have a particularly large proportion of these developed and agricultural lands, and the percentage of interior forest is lowest in the area (0 to 27 percent) (USFS 2011). The most affected lands are those on the fringes of major towns and cities, and in rural areas where second homes contribute to sprawling development (Irwin and Bockstael 2007, USFS 2011). Forest lands across the assessment area are often heavily dissected by roads, private property, trails, and utility lines. In Ohio, only 25 percent of the forest land is more than 0.25 mile from a road (Widmann et al. 2009). Parcelization is also a concern as the number of forest land owners is increasing and the size of parcels is decreasing (Widmann et al. 2009). Fragmentation of natural landscapes creates isolated plant and animal populations that are unable to migrate easily and exchange genetic information, leading to reduction in biological and genetic diversity (Riitters 2011). It also causes increased incidence of edges along forest boundaries (Sisk et al. 1997). Fragmentation has also resulted in the degradation of watersheds, loss of wildlife habitat, increased disturbances, and the spread of invasive species (Widmann et al. 2007). Natural Disturbances Natural disturbance has historically been a regular influence on forest ecosystems in the assessment area. Forest systems have distinct disturbance regimes, characterized in part by the soils, landforms, and vegetation (McNab et al. 2007). Small-scale canopy disturbances are often caused by drought, wind, ice, snow, flooding, landslides, insect outbreaks, intermediate-intensity fires, 38 and pathogens (NatureServe 2011). Larger scale canopy disturbances, potentially affecting entire stands and swaths of forest across the landscape, include tornadoes, hurricanes, high-intensity fires, periodic flooding along major river floodplains, and catastrophic insect and pathogen outbreaks. Annual spring floods along rivers and streams are also typical disturbance events, but hydrology has been modified by channelization, drainage tiles, dams, roads, and other anthropogenic activities that change soil or runoff characteristics (NatureServe 2011). Beaver historically affected floodplains along small streams by building dams that sometimes killed relatively large stands of trees and created temporary ponds and wetlands; beaver remain a small disturbance agent in the contemporary landscape. insect Pests and Diseases Insect and disease outbreaks have long influenced the structure of forest ecosystems in the Central Appalachians. Before European settlement and the introduction of nonnative species, outbreaks were caused by native insect species, including the spring hemlock looper and forest tent caterpillar. Recent outbreaks of another native species, the southern pine beetle, have occurred in the New Jersey Pine Barrens, and increasing populations in southeast Ohio warrant monitoring of this pest (NRCS 2014a). International trade and the inadvertent movement of nonnative invasive species from countries around the world have amplified the amount of exposure to, and impacts on, tree species of the Central Appalachians region. Gypsy moth is a serious pest and has caused huge losses of basal area in valuable red and white oak (MDNR 2010, ODNR 2010b). Beech bark disease has resulted in mortality of beech trees across millions of acres in the eastern United States, and has yet to invade the majority of the beech range (Morin et al. 2007). The hemlock woolly adelgid has threatened hundreds of thousands of eastern hemlocks with needle loss, followed by branch dieback, and eventually death (Jonas et al. 2012). The emerald ash borer has caused mortality ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE in all ash species, including white ash, black ash, and green ash, resulting in the loss of more than 50 million trees between 2002 and 2009 (Kovacs et al. 2010). The Asian longhorned beetle is not confirmed in the assessment area, but its arrival from adjacent areas would result in damage and mortality to many species including maples, buckeyes, birches, willows, and elms (Townsend Peterson and Scachetti-Pereira 2004). Diseases, such as chestnut blight fungus, have virtually eliminated American chestnut as a canopy tree although chestnut stump sprouts and saplings persist in the understory (Merkle et al. 2007). One or more species of fungus in the genus Hypoxylon can injure or kill trees weakened by other factors, such as drought, logging, and root disease (Anderson 1995). Nonnaive and Invasive Plants Nonnative plant species are a risk to forest ecosystems when they become invasive. These species affect forest ecosystems through direct competition for resources, alteration of fire or hydrologic conditions, disruption of natural succession and pollination, and other cascading influences (Frelich et al. 2012, Tu et al. 2001). Invasive plant species can be introduced into native ecosystems by the transportation of seed on vehicles or equipment, on the soles of shoes, in manure from domestic or wild animals, or by wind and water. The FIA program has monitored 25 invasive species in the eastern United States since 2007 (Fig. ). In West Virginia, it is estimated that 28 percent of plant species occurring in the wild are nonnative invasive species (Kurtz 2013). Kudzu, glossy buckthorn, bush honeysuckle, autumn olive, crown vetch, Japanese knotweed, Japanese stiltgrass, garlic mustard, ailanthus, mile-a-minute, and multiflora rose are among the area’s most problematic invasives (Grafton 2013). Figure 6.—Number of invasive plant species observed per FIA plot (2005 to 2010) (Kurtz 2013). 39 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Shits in Fire Regime Fire regimes have shifted in the assessment area over the past several hundred years, and these shifts inluence the composition of forest ecosystems. Both natural and human-caused ire has been a component of the eastern United States for thousands of years, although the return interval, intensity, and extent are largely dependent on landscape position in the Central Appalachians region (Abrams 1992, Nowacki and Abrams 2008, Thomas-Van Gundy and Nowacki 2013). Studies in the eastern United States and Canada that have dated ire scars in oak forests have shown that ire-return intervals either increased or decreased immediately after European settlement, depending on the stand and location. A 400-year reconstruction of ire history in western Maryland used ire scars as evidence that pre-European ires were as important as post-European ires, and that ire suppression in the 20th century has coincided with the increase of shade-tolerant species, to the detriment of oaks, hickories, and other ire-tolerant species (Shumway et al. 2001). The historic role of ire in the development and maintenance of oak forests has been well-established in the literature (Abrams 1992, Brose and Van Lear 1999). By the 1950s, ire exclusion began to favor red maple, sugar maple, American beech, and black cherry (Brose and Van Lear 1999, Nowacki and Abrams 2008, Schuler and Gillespie 2000, Wright and Bailey 1982). Oaks continue to be replaced by other hardwood species, especially red maple (Brose et al. 2008). Mineral, Gas, and Wind Energy Development Coal mining, natural gas fracturing, and wind power development are today’s most inluential natural resource extraction activities in the Central Appalachians. Coal mining is the dominant driver of land-use change in West Virginia, primarily changing forested conditions to nonforest (Liu et al. 2006). The mountaintop removal form of surface mining is a process that removes tons of bedrock from the sides and tops of mountains to reach the 40 underlying coal seams. The dramatic alteration of the landform is permanent, and waste disposal into valleys has resulted in the burial of headwater streams. Although the Surface Mining Control & Reclamation Act provides federal regulations on coal mining and reclamation operations, mined and reclaimed areas generally have lower iniltration capacity and higher runoff than pre-mine conditions (Townsend et al. 2009). Grading of the topsoil and subsoil, followed by seeding with grasses and herbs, has generally resulted in nearly impervious reclamation lands with compacted soils and herbaceous cover, rather than native forest (Bussler et al. 1984, Chong and Cowsert 1997, Holl 2002, Negley and Eshleman 2006, Simmons et al. 2008). Alterations to the water table, transition to overland low as the dominant runoff process, and increases in peak streamlow also are common consequences of mine land reclamation (Negley and Eshleman 2006). Natural gas wells irst appeared in the mid-1800s, and new wells continue to be drilled even as old wells are capped. In West Virginia and Ohio, there are more than 87,400 active gas-producing wells, with 487 horizontally drilled Marcellus wells (Kasey 2012, Resources for the Future 2012, WVDEP 2013a). Maryland had only seven gas wells operating in 2010, and although the Marcellus shale formation extends into western Maryland, horizontal drilling permits have been denied, pending research reports on the safety of hydraulic fracturing (O’Malley 2011). Electricity produced from natural gas creates approximately half the carbon dioxide emissions of electricity produced from coal; however, hydraulic fracturing requires extensive road and pipeline networks and millions of gallons of water. It is estimated that approximately 30 acres of land are disturbed for each shale gas drilling pad (Drohan et al. 2012b). Many well pads are located on soils with high to very high runoff potential, making actions to minimize long-term ecosystem degradation critical (Drohan et al. 2012a). ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Wind turbines in the Allegheny Mountains, West Virginia. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. Although wind energy has the potential to reduce greenhouse gas emissions and other adverse emissions, wind turbine placement has had notable ecological impacts, such as land surface disturbance (1 to 7 acres per turbine), road construction, vegetation clearing, soil removal and compaction, and fragmentation of forests (National Research Council [NRC] 2007). These disturbances can have subsequent impacts on forest composition and structure, as well as forest species sensitive to edge effects. Collision with wind turbines can cause significant mortality of birds and bats. Turbine design, site characteristics, location, and temporal patterns of use can all influence the rates of bird and bat mortality, but changes in how the turbines are operated can reduce mortality (NRC 2007). FOREST-DEPENDENT WILDLIFE The Central Appalachians region is one of the most ecologically diverse regions of the eastern United States (The Nature Conservancy 2003). The variations in topography, geology, and temperature and precipitation regimes in the region have resulted in the development of an exceptional variety of habitats supporting an abundance of wildlife, including more than 540 species identified as species of conservation concern in West Virginia alone (West Virginia Division of Natural Resources [WVDNR] 2014). 41 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE One of the most prevalent and well-known wildlife species is the white-tailed deer. It was almost eliminated from the region in the early 1900s as a result of deforestation and unregulated hunting, but now exists at higher densities than in the past several hundred years. Since the extirpation of the eastern cougar and eastern timber wolf from the region, deer have few natural predators to control population numbers, although black bear, bobcat, and coyotes do prey on fawns opportunistically. White-tailed deer, which can double in population size annually under optimum conditions, have exceeded their environmental carrying capacity in some areas (Côté 2004, Waller and Alverson 1997). At high densities, deer can have a keystone effect on the forest ecosystem. As deer browse plant species preferentially, they change the relative abundance and diversity of native species and promote the establishment of invasive species (Abrams and Johnson 2012, Collins and Carson 2003, Horsley et al. 2003, Knight et al. 2009). High deer densities can alter the availability of food to other wildlife species such as wild turkey and eastern gray squirrels that also rely on hard mast crops. Although deer occur throughout a variety of forest types in the assessment area, the distributions of some wildlife species are limited to specific White-tailed deer, which can have a signiicant impact on vegetaion near the forest loor. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 42 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE environmental conditions for all or a portion of their life cycle. Vernal pool obligate species, such as the spotted salamander, tiger salamander, and red-spotted newt breed only in isolated wetlands that are dry for part of the year. Because these ephemeral wetlands can occur where other water resources are scarce, they also provide important habitat for migratory and resident birds, large and small mammals (e.g., bats), and other species. Forestry practices also often result in destruction of these habitats because they are difficult to distinguish from the surrounding forest when pools have dried up in the summer. The spruce/fir forest ecosystem of the Allegheny Mountains section provides other unique habitats supporting endemic and obligate species. Many sensitive wildlife species residing there are competitive only in the microclimates provided by high-elevation and complex topography. The Cheat Mountain salamander is listed as a federally threatened species and the West Virginia northern flying squirrel was removed from the endangered species list in 2013, but remains a USFS Regional Forester’s Sensitive Species and a State Species of Greatest Conservation Need in West Virginia (WVDNR 2014). The dense shading and moist microclimate associated with spruce and spruce/ northern hardwood forests, along with highly organic and often acidic and rocky soils beneath the conifers, provide a habitat where the Cheat Mountain salamander may have a competitive advantage over other salamanders, such as the red-backed salamander, which is dominant at lower elevations. The northern flying squirrel also has an intricate relationship with these habitats. Like the Cheat Mountain salamander, the northern flying squirrel competes with a similar species, the southern flying squirrel, which is dominant at lower elevations. Many boreal bird species are characteristic of these high-elevation habitats as well, ranging from predatory birds such as the northern goshawk and saw-whet owl to Neotropical Ephemeral pool. Seasonally wet areas like this one are important for amphibian reproducion. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. migrants such as the blackburnian warbler and red crossbill, all of which reach the southeastern extent of their breeding ranges in the Central Appalachians. Deciduous forests in the region also provide critical breeding bird habitat. Four of the birds of highest conservation priority in the Appalachian Mountains region are hardwood interior forest species: the cerulean warbler, Kentucky warbler, wood thrush, and worm-eating warbler (Appalachian Mountains Joint Venture Board 2008). Within the assessment area, these species require mature deciduous or mixed forest habitat for breeding, and each is associated with particular herbaceous and understory structure. These Neotropical migrants are vulnerable to a variety of threats, including tropical deforestation on wintering grounds and forest habitat loss, fragmentation, and modification of breeding grounds. Another bird common in mature oak and oak-pine forests in the region is the wild turkey, an important game species (USFWS 2010). Bats are also associated with forested habitats and are primary predators of nocturnal insects, including many forest and agricultural pests. Populations of many bat species in the eastern United States are 43 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE in a state of rapid decline as a result of white-nose syndrome, first detected in New York in 2007. This disease is caused by the fungus Geomyces destructans, which has spread through hibernacula throughout eastern North America. Bat mortality can reach 100 percent at infected sites. As of 2012, approximately  million bats had died from the disease in the United States and Canada. The syndrome was confirmed in West Virginia in the winter of 2008-2009 and in Maryland and Ohio in the following two winters. It has already had a major impact on many bat species, including the endangered Indiana bat, northern bat, little brown bat, small-footed bat, and tri-colored bat. The last four “forest bats” are on the USFS Regional Forester’s Sensitive Species list as well as state sensitive species listings, and may be soon proposed for federal listing. Riparian habitats can be critical for many other wildlife species, providing breeding and foraging habitat for a wide variety of waterfowl, amphibians, reptiles, and mammals, as well as diverse invertebrate fauna. The beaver, another species that was nearly extirpated at the start of the 20th century, requires riparian systems for habitat and plays a keystone role in creating and maintaining open water wetland habitats. The brook trout is the only trout species native to the assessment area and much of the eastern United States, and is often used as an indicator of the health of a watershed. Primary threats to brook trout include poor land management, high water temperatures, urbanization, acid deposition and runoff, sedimentation, surface and ground water withdrawals and impoundments, and introduction of nonnative fish species. As a result of these and other factors, brook trout populations have been greatly reduced in Ohio and West Virginia, and only three intact subwatersheds remain in the western panhandle of Maryland (Trout Unlimited 200). 44 CuRRENT LAND MANAGEMENT TRENDS Forest ownership There are numerous types of forest landowners within the assessment area (Table 7, Fig. 7). About 14 percent of forest land in the region is publicly owned. National forests and state land compose the largest percentages of public forest land, followed by land owned by county and municipal governments, the National Park Service, and the U.S. Department of Defense. The Monongahela National Forest administers approximately 920,000 acres in West Virginia, and the Wayne National Forest administers approximately 250,000 acres in Ohio. Most of the forests in the assessment area, however, are privately owned. This category reflects a diversity of landowner types, including industrial and corporate organizations, conservation organizations, families, and individuals. As a result, private ownership patterns are complex and change over time. Trends in Forest use and Management Most private forest land is held by hundreds of thousands of nonindustrial family forest owners Table 7.—ownership categories of forest land in the assessment area (USFS 2013) Ownership Forest land (acres) % Private Naional forest State County and municipal Naional Park Service Department of Defense Other federal Other local government U.S. Fish and Wildlife Service 16,194,332 1,267,057 946,630 251,036 85,911 80,562 34,622 22,553 11,826 85.7 6.7 5.0 1.3 0.5 0.4 0.2 0.1 0.1 Total 18,894,530 100 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Figure 7.—Public and private forest ownership within the assessment area (Hewes et al. 2014). (Butler 2008). The primary reasons for forest ownership are privacy, scenery, part of home or cabin, nature protection, to pass land on to heirs, and for access to hunting and fishing (Butler 2008). A survey of 5-year management plans identified a variety of landowner goals and management styles: minimal or no activity, harvesting firewood, harvesting pulp and sawlogs, transferring to heirs, and buying more forest land. Family owners can enroll their lands in conservation easements or forest certification programs such as the American Tree Farm System (ATFS), which require forests to have written management plans (Box 3). About 1.5 million acres in Ohio, West Virginia, and Maryland are currently certified by the ATFS (ATFS 2014). Engaged family forest owners often look to extension agents, Conservation Districts, and private consultants to provide technical assistance and other resources for managing forests. Industrial forest landowners manage forest lands for timber products, and have a vested interest in long-term forest management. Millions of acres of corporate land have been transferred in the last decade to REITs and TIMOs, which are considered private nonindustrial forest landowners, largely due to unfavorable taxation on industry-owned forests (Froese et al. 2007, Zhang et al. 2012). REITs own and operate income-producing real estate and timberland holdings, sometimes made public through trading of shares on a stock exchange, and thus are able to take advantage of more favorable tax policies. TIMOs act as investment managers for institutional clients who own the timberlands as investments or partnership shares (Fernholz et al. 2007). The goal in both cases is to maximize the growth of the timberland asset over the short term. Thus, the purchase of timberland by REITs and TIMOs raises concerns about parcelization, 45 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE development, and high-yield management practices (Fernholz et al. 2007). TIMOs in particular have managed forested lands similar to high-intensity industrial forests, with a high percentage in pine plantations (Fernholz et al. 2007). Public (federal, state, and county) agencies and tribal organizations own extensive tracts of forest in the assessment area. These lands are often managed to provide many environmental benefits, often including wildlife habitat, water protection, soil conservation, nature preservation, timber production, recreation, cultural resources, and a variety of other uses (MDNR 2010; ODNR 2010b; USFS 200a, 200b). Box 3: Programs for Private Landowners All three states ofer incenives to private forest landowners, with the intent of maintaining larger parcels of privately owned forest and promoing sustainable producion of forest products. About 85 percent of the forested land in the area is privately owned, however, and most of these lands lack a management plan (USFS 2008). ohio More than 65,000 acres of private forest lands in Ohio are enrolled in the Ohio Forest Tax Law program under the “new law” rules implemented in 1993 with the overarching goal to protect land from urban sprawl. This program requires at least 10 acres and a commitment to manage for soil and water conservaion and producive forest land. Together with protected lands, about 870,000 acres, or roughly 10 percent of Ohio’s forests, have commitments to soil and water conservaion (ODNR 2010b). The Current Agricultural Use Value program is designed to promote imber producion, and property assessment values are reduced to $100 per acre. To qualify, a landowner must devote land exclusively to agricultural use, which includes the growth of imber for a noncommercial purpose. This program does not require a management plan (ODNR 2014). Maryland Maryland also administers programs to help ease property taxes and maintain healthy forests. The Forest Conservaion Management Agreement 4 (FCMA) is a conservaion easement program that lowers assessed values to $125 per acre on a minimum of 5 acres and a minimum expiraion date of 15 years (MDNR 2014). As of January 2014, 1,300 landowners had 84,000 acres enrolled. In the Western Region, which overlaps the Maryland porion of the assessment area, 234 landowners had 17,670 acres enrolled (Tim Culbreth, MDNR, pers. commun.). The Woodland Assessment Program is a county program; similar to FCMA, there are no enrollment fees or imeframe, but property assessment values are reduced to $187.50 per acre of forest. The Maryland Income Tax Modiicaion program allows woodland owners to deduct double the cost for reforestaion and imber stand improvement pracices on 3 to 1,000 acres from the federal adjusted gross income on the Maryland tax return. West Virginia In West Virginia, the Managed Timberland Program provides tax incenives for forest landowners who pracice sustainable forestry on their nonindustrial, privately owned forestland comprising 10 acres or more (West Virginia Division of Forestry 2012). Paricipaion in this program has been growing steadily since 1997, and there are now nearly 2.4 million acres enrolled. Many paricipaing landowners have also paricipated in the USFS Forest Stewardship Program in order to receive assistance with wriing a forest management plan at a reduced cost (Dye 2013). ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE Forest Ceriicaion Forest certification is a process designed to ensure that forest products originate from forests that are sustainably managed. Forest lands are certified through several systems, including the Forest Stewardship Council (FSC), the Sustainable Forestry Initiative (SFI), and the ATFS (Table 8). The Ohio Department of Natural Resources has dual certification for sustainable forest management of its state forests through FSC and SFI, with a total of 202,927 certified acres. Nearly all of Maryland’s 211,000 acres of state forests are dual certified under FSC and SFI. Timber Harvest and Forest Products As mentioned above, the forestry sector is a notable economic contributor in the assessment area. Within the Central Appalachians region, forest removals (not including mortality) averaged 277 million cubic feet in 2011 (USFS 2013). Over half the total harvested roundwood was used as pulpwood and around 30 percent was used as saw logs, with the remainder being diverted to a variety of uses or left behind as logging slash. Pulpwood production peaked in 1994 and has been gradually declining over recent years. More specific harvest data is available at the state level (Table 9). Across the assessment area, hardwoods account for most commercial species, including tulip tree, red and white oaks, soft and hard maples, and black cherry (Piva and Cook 2011, Walters et al. 2008, Wiedenbeck and Sabula 2008). In West Virginia, 97 percent of industrial roundwood processed in 2007 consisted of hardwood species, 37 percent of which was tulip tree (Piva and Cook 2011). The FIA data also provide more information about the amount of wood removed from forests in the assessment area through timber harvest or conversion of forest to nonforest, with the vast Table 8.—Forest land enrolled in ceriicaion programs (acres)a State Forest Stewardship Council Council (FSC) Maryland Ohio West Virginia a – 203,957 39,039 Forest land enrolled in ceriicaion program (acres) Sustainable Forestry American Tree Farm Iniiaive (SFI) System (ATFS) – 202,927 257,044 139,021 293,585 1,013,352 Dual-ceriied (FSC and SFI) 211,000 202,927 – Data compiled from muliple sources (ATFS 2014, Forest2Market 2013, SFI 2013). Table 9.—Statewide average annual roundwood removals in million cubic feet (Piva and Cook 2011, Walters et al. 2008, Wiedenbeck and Sabula 2008) Average annual roundwood removals (milllion cubic feet) Ohio (2003 to 2006) Maryland (2008) West Virginia (2007) Total industrial roundwood Domesic logs Pulpwood 91.2 67.7 23.5 29.1 16 12.5 189 104 66.7 47 ChAPTER 1: ThE CoNTEMPoRARY LANDSCAPE majority of removals in this region being due to timber harvest. The amount of wood harvested annually in the Central Appalachians region is less than the amount that is grown each year, suggesting that the harvest of timber products is biologically sustainable (Lister and Perdue 2013, Widmann and Morin 2012, Widmann et al. 2007). The net annual growth-to-removal ratio is based upon FIA data and provides a primary measure of sustainability. This ratio compares net growth (i.e., gross growth minus mortality) to removals from forest management for forested lands; values greater than 1.0 indicate that net annual growth is greater than annual removals and that the removal rate is sustainable. Across all ownership classes in the assessment area, the growth-to-removal ratio was 2.3 for the most recent inventory period (2008 through 2012), meaning that growth was more than double removals (Table 10). Among ownership classes in the assessment area, national forests and national parks have the highest growth-to-removal ratio, indicating low levels of harvest compared to other owners (Table 10). ChAPTER SuMMARY The climate, geology, and soils of the Central Appalachians region of Ohio, West Virginia, and Maryland support a mosaic of forest ecosystems. These communities supply important benefits to the people of the area, including forest products and recreation opportunities. Past changes in climate, fire regime, and land use have shaped the landscape into its current condition. Shifts in fire regime, habitat fragmentation, species invasions, insect pests and diseases, and other alterations to the landscape threaten the integrity and diversity of the ecosystems and the benefits they provide. Management on public lands in recent decades has focused on reducing these stressors and improving ecosystem function. About 85 percent of the forested land in the area is privately owned, however, and the majority of these lands lack a management plan. New opportunities and incentives have arisen in recent years to help private and public land managers to restore and conserve the ecosystems of the Central Appalachians for future generations. Table 10.—Growth, mortality, and removals of growing stock on forest land in the assessment area (USFS 2013) Ownership Annual net growth (cubic feet) Annual removals (cubic feet) Annual net growth:removals Naional Forest Naional Park Service U.S. Fish and Wildlife Service Department of Defense Other federal State County and municipal Other local government Private Othera 42,954,113 5,706,942 508,390 1,967,529 757,053 28,140,927 10,964,142 1,403,549 768,866,740 2,234,386 324,147 66,516 – – – 10,821,876 202,251 258,295 310,059,634 52,090,765 132.5 85.8 – – – 2.6 54.2 5.4 2.5 0.0 Total 863,503,771 373,823,484 2.3 a Represents esimated net growth and removals for lands that were diverted from forest and nonforest. 48 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG This chapter provides a brief background on climate change science, climate simulation models, and models that project the impacts of changes in climate on tree species and ecosystems. Throughout the chapter, boxes indicate resources to find more information on each topic. The resources listed are up-to-date, nontechnical reports based on the best available science. A more detailed scientific review of climate change science, trends, and modeling can be found in the Intergovernmental Panel on Climate Change (IPCC) Fourth Assessment Report (IPCC 2007), and the third National Climate Assessment (Melillo et al. 2014). CLiMATE ChANGE Climate is not the same thing as weather. Weather is a set of the meteorological conditions for a given point in time in one particular place (such as the temperature at 3:00 p.m. on June 22 in Athens, OH). Climate, in contrast, is the long-term average of meteorological conditions and patterns for a geographic area. This climate average is calculated from individual measurements taken at multiple locations across a geographic area, and at different points through time. The IPCC (2007: 30) defines climate change as “a change in the state of the climate that can be identified (e.g., by using statistical tests) by changes in the mean and/or the variability of its properties, and that persists for an extended period, typically decades or longer.” A key finding of the IPCC in its Fourth Assessment Report (2007) was that “warming of the climate system is unequivocal.” This was the first assessment report in which the IPCC considered the evidence strong enough to make such a statement. Current observations of higher global surface, air, and ocean temperatures and thousands of long-term (more than 20 years) data sets from all continents and oceans contributed to this conclusion. These data sets showed significant changes in snow, ice, and frozen ground; hydrology; coastal processes; and terrestrial, marine, and biological systems. The IPCC’s Fifth Assessment Report contains the most recent and comprehensive evidence of global changes synthesized to date (see Box 4 for a link to the draft). Selected global and national assessments are listed in Box 4. The Warming Trend The Earth is warming, and the rate of warming is increasing (IPCC 2007). Measurements from weather stations across the globe indicate that the global mean temperature has risen steadily over the past 50 years, and that the year 2011 was 0.9 °F (0.5 °C) warmer than the 1951 to 1980 mean (IPCC 2007) (Fig. 8). The first 13 years of the 21st century rank among the warmest 14 years in the 134-year period of record of global temperature (National Oceanic and Atmospheric Administration [NOAA] 2014b). Temperatures in the United States have risen by 2 °F (1.1 °C) in the last 50 years (Karl et al. 2009). The 2012 continental U.S. average annual temperature of 55.3 °F was 3.1 °F above the 20th-century average, and was the warmest year in the 1895 to 2013 period of record for the nation (NOAA 2014b). 49 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Box 4: Global and Naional Assessments Intergovernmental Panel on Climate Change u.S. Global Change Research Program The Intergovernmental Panel on Climate Change (IPCC; htp://www.ipcc.ch/) is the leading internaional body for the assessment of climate change. It was established by the United Naions Environment Programme (UNEP) and the World Meteorological Organizaion (WMO) in 1988 to provide the world with a clear scieniic view on the current state of knowledge in climate change and its potenial environmental and socioeconomic impacts. Its Fith Assessment Report consists of the Climate Change 2014 Synthesis Report and reports by Working Groups I, II, and III. Drats of these reports are available for download at the Web address below. Please note that Web addresses are current as of the publicaion date of this assessment but are subject to change. The U.S. Global Change Research Program (USGCRP; htp://www.globalchange.gov/) is a federal program that coordinates and integrates global change research across 13 government agencies to ensure that it efecively and eiciently serves the naion and the world. Mandated by Congress in the Global Change Research Act of 1990, the USGCRP has since made the world’s largest scieniic investment in the areas of climate science and global change research. It has released several naional synthesis reports on climate change in the United States, which are available for download at the Web addresses below. Climate Change 2014: Synthesis Report and Working Group contribuions to the Fith Assessment Report www.ipcc.ch/report/ar5/ Climate Change 2007: Synthesis Report www.ipcc.ch/publicaions_and_data/ar4/syr/en/ contents.html Synthesis and Assessment Products htp://library.globalchange.gov/products/ assessments/ Naional Climate Assessment htp://nca2014.globalchange.gov/ Efects of Climaic Variability and Change on Forest Ecosystems: a Comprehensive Science Synthesis for the u.S. www.treesearch.fs.fed.us/pubs/42610 Figure 8.—Trends in global temperature compared to the 1951 to 1980 mean. Data source: NASA Goddard Insitute for Space Studies. Image courtesy of NASA Earth Observatory, Robert Simmon; www.giss.nasa.gov/research/news/20120119/. 50 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Average annual global temperature increases of the last 50 years are just one aspect of a more complex and wide-ranging set of climatic changes. For example, the frequency of cold days, cold nights, and frosts has decreased over many regions of the world while the frequency of hot days and nights has increased (IPCC 2007). The frequency of heat waves and heavy precipitation events has increased over this period, with new records for both heat and precipitation in portions of the United States in July 2011 and March 2012 (NOAA 2012). Global rises in sea level, decreasing extent of snow and ice, and shrinking of mountain glaciers have all been observed over the past 50 years, and are consistent with a warming climate (IPCC 2007). Average temperature increases of a few degrees may seem small, but even small increases can result in substantial changes in the severity of storms, the nature and timing of precipitation, droughts and heat waves, ocean temperature and volume, and snow and ice—all of which affect humans and ecosystems. Temperature increases above 3. °F (2 °C) are likely to cause major societal and environmental disruptions through the rest of the century and beyond (Richardson et al. 2009). The synthesis report of the International Scientific Congress on Climate Change concluded that “recent observations show that societies and ecosystems are highly vulnerable to even modest levels of climate change, with poor nations and communities, ecosystem services and biodiversity particularly at risk” (Richardson et al. 2009: 12). Based on available evidence, 97 percent of the climate science community attributes this increase in temperature and associated changes in precipitation and other weather events to human activities (Anderegg et al. 2010, Cook et al. 2013, Doran and Zimmerman 2009, Stott et al. 2010). Scientists have been able to attribute these changes to human causes by using climate model simulations of the past, both with and without human-induced changes in the atmosphere, and then comparing those simulations to observational data. Overall, these studies have shown a clear human “fingerprint” on recent changes in temperature, precipitation, and other climate variables due to changes in greenhouse gases and particulate matter in the air (Stott et al. 2010). Chapter 3 provides specific information about recent climate trends for the assessment area. The Greenhouse Efect The greenhouse effect is the process by which certain gases in the atmosphere absorb and re-emit energy that would otherwise be lost into space (Fig. 9). The greenhouse effect is necessary for human survival: without it, Earth would have an average temperature of about 0 °F (-18 °C) and be covered in ice, rather than a comfortable 59 °F (15 °C). Several naturally occurring greenhouse gases in the atmosphere, including carbon dioxide (CO2 ), methane, nitrous oxide, and water vapor, contribute to the greenhouse effect. Water vapor is the most abundant greenhouse gas; its residence time in the atmosphere, however, is on the order of days as it responds to changes in temperature and other factors. Carbon dioxide, methane, nitrous oxide, and other greenhouse gases reside in the atmosphere for decades to centuries. Thus, these other long-lived gases are of primary concern with respect to longterm warming. Human Inluences on Greenhouse Gases Humans have increased the concentrations of carbon dioxide, methane, nitrous oxide, and halocarbons in the atmosphere since the beginning of the industrial era (Fig. 10). More carbon dioxide has been released by humans into the atmosphere than any other greenhouse gas. Carbon dioxide levels increased at a rate of 1.4 parts per million (ppm) per year from 190 to 2005 (IPCC 2007), and reached an average of 395 ppm in January 2013 (Tans and Keeling 2013). In recent decades, fossil fuel burning has accounted for an estimated 83 to 94 percent 51 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Figure 9.—Idealized model of the natural greenhouse efect. Figure courtesy of IPCC (2007). Figure 10.—Concentraions of greenhouse gases showing increases in concentraions since 1750 atributable to human aciviies in the industrial era. Concentraion units are parts per million (ppm) or parts per billion (ppb), indicaing the number of molecules of the greenhouse gas per million or billion molecules of air. Figure courtesy of IPCC (2007). 52 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG of the human-induced increase in carbon dioxide. The remaining  to 17 percent of human-induced emissions comes primarily from deforestation of land for conversion to agriculture, which releases carbon dioxide when forests burn or decompose (van der Werf et al. 2009). However, increases in fossil fuel emissions over the past decade mean that the contribution from land-use changes has become a smaller proportion of the total (Le Quéré et al. 2009). Methane is responsible for roughly 14 percent of greenhouse gas emissions in terms of CO2equivalent (CO2-eq) (IPCC 2007). Concentrations of this gas have also been increasing as a result of human activities, including agricultural production of livestock and increases in rice production. Livestock production contributes to methane emissions primarily from fermentation in the guts of cattle and other ruminants. Rice production requires wet conditions that are also ideal for microbial methane production. Other sources of methane include biomass burning, microbial-induced methane emissions from landfills, fossil fuel combustion, and leakage of natural gas during extraction and distribution. Nitrous oxide accounts for about 8 percent of global greenhouse gas emissions in terms of CO2-eq (IPCC 2007). The primary human source of nitrous oxide is agriculture. The use of fertilizer causes emissions from soil as microbes break down nitrogencontaining products. This is especially dramatic in areas where tropical forests are converted to agricultural lands. Other human-caused sources of nitrous oxide include nylon production and combustion of fossil fuels. Humans have also reduced ozone, which protects us from ultraviolet radiation, in the atmosphere through the use of chlorofluorocarbons (CFCs) once used widely in refrigeration, air conditioning, and other uses. Restrictions against the use of CFCs under the Montreal Protocol led to a decline in CFC emissions, and reductions in ozone have subsequently slowed. After CFCs were banned, another class of halocarbons, hydrofluorocarbons (HFCs, also known as F-gases), largely replaced CFCs in refrigeration and air conditioning. HFCs do not deplete stratospheric ozone, but many are powerful greenhouse gases. Currently HFCs account for about 1 percent of greenhouse gas emissions in terms of CO2-eq (IPCC 2007). CLiMATE MoDELS Scientists use models, which are simplified representations of reality, to simulate future climates. Models can be theoretical, mathematical, conceptual, or physical. General circulation models (GCMs) combine complex mathematical formulas representing physical processes in the ocean, atmosphere, and land surface within large computer simulations. In this assessment, GCMs are used to project future climate and as inputs to impact models. General Circulaion Models General circulation models simulate physical processes in the earth, oceans, and atmosphere through time using mathematical equations in threedimensional space. They can work in time steps as small as minutes or hours in simulations covering decades to centuries. Because of their high level of complexity, GCMs require intensive computing power, and must be run on supercomputers. Although climate models use highly sophisticated computers, limits on computing power mean that projections are limited to relatively coarse spatial scales. Instead of simulating climate for every single 53 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG point on Earth, modelers divide the land surface, ocean, and atmosphere into a three-dimensional grid (Fig. 11). Each cell within the grid is treated as an individual unit, and is able to interact with adjacent cells. Although each model is slightly different, the size of each cell in the grid is usually between 2 and 3° latitude and longitude, or for the middle latitudes, about the size of West Virginia. These horizontal grids are stacked in interconnected vertical layers that simulate ocean depth or atmospheric thickness at increments usually ranging from 50 to 3,280 feet. Figure 11.—Schemaic describing climate models, which are systems of diferenial equaions based on the basic laws of physics, luid moion, and chemistry. The planet is divided into a three-dimensional grid that is used to apply basic equaions; atmospheric models calculate winds, heat transfer, radiaion, relaive humidity, and surface hydrology within each grid and evaluate interacions with neighboring points. Figure courtesy of NOAA (2008). 54 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Several research groups from the United States and abroad have developed GCMs that have been used in climate projections for the IPCC reports and elsewhere (Box 5). These models have been developed by internationally renowned climate research centers such as NOAA’s Geophysical Fluid Dynamics Laboratory (GFDL CM2) (Delworth et al. 200), the United Kingdom’s Hadley Centre (HadCM3) (Pope et al. 2000), and the National Center for Atmospheric Research (PCM) (Washington et al. 2000). These models use slightly different grid sizes and ways of quantitatively representing physical processes. They also differ in sensitivity to changes in greenhouse gas concentrations, which means that some models will tend to project higher increases in temperature than others under the same greenhouse gas concentrations (Winkler et al. 2012). Like all models, GCMs have strengths and weaknesses (Box ). In general, they are useful and reliable tools because they are based on wellunderstood physical processes and have been judged in part by their ability to accurately simulate past climate. Simulations with GCMs can be run for past climate, and output from these simulations generally correspond well with proxy-based estimates of ancient climates and actual historical measurements of recent climates. Projections by GCMs are not perfect, however. Sources of error in model output include incomplete scientific understanding of some climate processes and the fact that some influential climate processes occur at spatial scales that are too small to be modeled with current computing power. Technological advances in the computing industry along with scientific advances in our understanding of Earth’s physical processes will lead to continued improvements in GCM projections. Emissions Scenarios General circulation models require significant amounts of information to project future climates. Some of this information, like future greenhouse gas concentrations, is not known and must be estimated. Although human populations, economies, and technological developments will certainly affect future greenhouse gas concentrations, these developments cannot be completely foreseen. One common approach for dealing with uncertainty about future greenhouse gas concentrations is to develop storylines (narratives) about how the future may unfold and calculate the potential greenhouse gas concentrations for each storyline. The IPCC’s set of standard emissions scenarios is a widely accepted set of such storylines (IPCC 2007). In GCMs, the use of different emissions scenarios results in different climate projections. Box 5: More Resources on Climate Models and Emissions Scenarios U.S. Forest Service Intergovernmental Panel on Climate Change Climate Projecions FAQ www.treesearch.fs.fed.us/pubs/40614 Chapter 8: Climate Models and Their Evaluaion www.ipcc.ch/publicaions_and_data/ar4/wg1/en/ ch8.html u.S. Global Change Research Program Special Report on Emissions Scenarios: Summary for Policymakers htp://www.ipcc.ch/ipccreports/sres/emission/index. php?idp=0 Climate Models: an Assessment of Strengths and Limitaions library.globalchange.gov/sap-3-1-climate-modelsan-assessment-of-strengths-and-limitaions 55 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Box 6: Model Limitaions and Uncertainty “All models are wrong, some are useful.” –George Box (Box and Draper 1987) Models are conceptual representaions of reality, and any model output must be evaluated for its accuracy to simulate a biological or physical response or process. The overall intenion is to provide the best informaion possible for land managers given the uncertainty and limitaions inherent in models. Model results are not considered standalone components of this vulnerability assessment because there are many assumpions made about the processes simulated by GCMs and impact models, uncertainty in future greenhouse gas concentraions, and limitaions on the grid scale and numbers of inputs that a model can reliably handle. Precipitaion projecions usually have much more variability among models than temperature. Regions with complex topography contain much more diversity in microclimates than many models can capture. Many nonclimate stressors, such as insect pests or pathogens, can overshadow the impact of climate on a species or community, especially in the short term. Therefore, model results used in this assessment were evaluated by local experts to idenify regional caveats and limitaions of each model, and are considered with addiional knowledge and experience in the forest ecosystems being assessed. Emissions scenarios quantify the effects of alternative demographic, technological, or environmental developments on atmospheric greenhouse gas concentrations. None of the current scenarios includes any changes in national or international policies, such as the Kyoto Protocol, directed specifically at climate change. However, some of the scenarios that include a reduction in greenhouse gases through other means suggest what we could expect if these policies were implemented. Six different emissions scenarios are commonly used in model projections for reports such as the IPCC Fourth Assessment Report (Fig. 12). 5 We integrated fundamentally diferent types of impact models into our assessment of forest vulnerability to climate change. These models operate at diferent spaial scales and provide diferent kinds of informaion. The DISTRIB model projects the amount of available suitable habitat for a species. The LINKAGES model projects species establishment and growth. The LANDIS PRO model projects changes in basal area and abundance. There are similariies between some inputs into these models—downscaled climate models and scenarios, simulaion ime periods, and many of the same species—but because of the fundamental diferences in their architecture, their results are not directly comparable. Their value lies in their ability to provide insights into how various interrelated forest components may respond to climate change under a range of possible future climates. Models can be useful, but they are inherently incomplete. For that reason, an integrated approach using muliple models and expert judgment is needed. The basic inputs, outputs, and architecture of each model are summarized in this chapter with clear descripions of the limitaions and caveats of each model. Limitaions of these models with speciic applicability to the Central Appalachians forest ecosystems are discussed in more detail in Chapter 5. The A1FI scenario is the most fossil-fuel intensive, and thus projects the highest future greenhouse gas concentrations; GCM simulations using the A1FI scenario project the highest future warming. On the other end of the spectrum, the B1 scenario represents a future where alternative energies decrease our reliance on fossil fuels and greenhouse gas concentrations increase the least. GCM simulations using the B1 scenario project the lowest increase in global temperature. Although these scenarios were designed to describe a range of future emissions over the coming decades, it is important to note that the future will likely be different from any of the ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Figure 12.—Projected global greenhouse gas (GHG) emissions (in gigatons [Gt] of CO2-eq per year) assuming no change in climate policies under six scenarios (B1, A1T, B2, A1B, A2, and A1FI) originally published in the Special Report on Emissions Scenarios (SRES) (IPCC 2000), and the 80th-percenile range (gray shaded area) of recent scenarios published since SRES. Dashed lines show the full range of post-SRES scenarios. Figure courtesy of IPCC (2007). developed scenarios. It is highly unlikely that future greenhouse gas emissions will be less than described by the B1 scenario even if national or international policies were implemented immediately. In fact, current emissions more closely track the greenhouse gas emissions of the A1FI scenario, and global emissions since the year 2000 have even exceeded those values in some years (Raupach et al. 2007). Downscaling As mentioned previously, GCMs simulate climate conditions only for relatively large areas on a relatively coarse scale. To better examine the future climate of areas within the Central Appalachians region, a smaller grid scale is needed. One method of improving the resolution uses statistical downscaling, a technique by which statistical relationships between GCM model outputs and on-the-ground measurements are derived for the past (Hayhoe et al. 2007, Stoner et al. 2013). First, a statistical relationship is developed between GCM output for a past “training period,” and observed climate variables of interest (e.g., temperature and precipitation). The historical relationship between GCM output and monthly or daily climate variables at the regional scale can then be tested by using an independent historical “evaluation period” to confirm the relationship is robust. Finally, the historical relationship between GCM output and observed climate variables is used to downscale both historical and future GCM simulations to the same scale as the initial observations. The statistical relationships are then used to adjust large-scale GCM simulations of the future to much smaller spatial scales. The grid resolution for downscaled climate projections is typically around .2 miles (i.e., a cell represents 38.4 square miles). Statistical downscaling has several advantages and disadvantages (Daniels et al. 2012). It is a relatively simple and inexpensive way to produce smallerscale projections using GCMs. One limitation is that downscaling assumes that past relationships between large-scale weather systems and local climate will remain consistent under future change. Evaluation of this assumption was performed by applying the asynchronous regional regression model (ARRM) to a high-resolution (15.5 miles) GCM data set under the new RCP 8.5 scenario, and comparing the high-resolution output directly to the projections using SRES scenarios (Hayhoe et al. 2013). The RCP 8.5 scenario is one of the newest suite of scenarios developed by the IPCC, and is comparable to the SRES A1FI scenario used in this assessment. Hayhoe and others (2013) found that the assumption holds true for small projections of change, but larger projections of change may result in small biases. Maximum daily temperatures showed bias within the assessment area only for hot days, whereas minimum daily temperatures showed more widespread bias, indicating potential overestimation of increases in warm nights. 57 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Precipitation projections appear to have widespread bias within the assessment area only on the wettest days. Another limitation is that downscaling depends on local climatological data. If there are too few weather stations in the area of interest, it will be difficult to obtain a good downscaled estimate of future climate for that area. Finally, local influences on climate that occur at finer scales (such as land cover type or topography) cannot be addressed by statistical downscaling, adding to uncertainty when downscaling climate projections. Another approach, dynamical downscaling, uses a regional climate model (RCM) embedded within a GCM (Daniels et al. 2012). Like GCMs, RCMs simulate physical processes through mathematical representations on a grid. However, RCMs operate on a finer resolution than GCMs, typically ranging from 15.5 to 31 miles, but can be finer than .2 miles. Thus, they can more realistically simulate the effects of topography, land cover, lakes, and regional circulation patterns that operate on smaller scales. However, dynamical downscaling requires even more computational power than statistical downscaling. This means that dynamically downscaled data are usually available for only one or two GCMs or scenarios, and for limited geographic areas. Because dynamically downscaled data are currently limited for the assessment area, we use statistically downscaled data in this report. Downscaled GCMs used in this Assessment In this assessment, we report statistically downscaled climate projections for two modelemissions scenario combinations: GFDL A1FI and PCM B1 (unless otherwise noted). Both models and both scenarios were included in the IPCC Fourth Assessment Report (IPCC 2007). The latest version of the National Climate Assessment (NCA) (Melillo et al. 2014) also draws on statistically downscaled data based on IPCC models and scenarios but uses the A2 scenario as an upper bound, which projects 58 lower emissions compared to A1FI. The IPCC Assessment includes several other models, which are represented as a multi-model average in its reports. The NCA takes a similar approach in using a multi-model average. For this assessment, we instead selected two models that simulated climate in the eastern United States with low error and that bracketed a range of temperature and precipitation futures (Hayhoe 2010a). This approach gives readers a better understanding of the level of agreement among models and provides a range of alternative scenarios that can be used by managers in planning and decisionmaking. Working with a range of plausible futures helps managers avoid placing false confidence in a single scenario given uncertainty in projecting future climate. The GFDL model developed by NOAA is considered moderately sensitive to changes in greenhouse gas concentrations (Delworth et al. 200). In other words, any change in greenhouse gas concentration would lead to a change in temperature that is higher in some models and lower than others. The A1FI scenario is the highest greenhouse gas emissions scenario used in the 2007 IPCC assessment, and is most similar to current trends in global greenhouse gas emissions. Therefore the GFDL A1FI scenario represents a higher-end projection for future temperature increases. The PCM, in contrast, is considered to have low sensitivity to greenhouse gas concentrations. The B1 scenario is the lowest greenhouse gas emissions scenario used in the 2007 IPCC assessment, and is much lower than the most likely trajectory for greenhouse gas emissions for the coming decades. Therefore, the PCM B1 combination represents a lower-end projection of future climate change. Together, the GFDL A1FI and PCM B1 scenarios span a large range of possible future climate scenarios. Although both projections are possible, the GFDL A1FI scenario represents a more realistic ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG A woods road in the Scioto Trail State Park, Ohio. Photo by the Ohio Department of Natural Resources, used with permission. projection of future greenhouse gas emissions and temperature increases (Raupach et al. 2007). No likelihood has been attached to any of the emissions scenarios, however, and it is possible that actual emissions and temperature increases could be lower or higher than these projections (IPCC 2007). This assessment uses a statistically downscaled climate data set (Hayhoe 2010a). Daily mean, minimum, and maximum temperature and total daily precipitation were downscaled to an approximately 7.5-mile resolution grid across the United States. This data set uses a modified statistical ARRM to downscale daily GCM output and historical climate data (Stoner et al. 2013). Asynchronous quantile regression used historical gridded meteorological data from 190 through 1999 at 1/8-degree resolution (.2 to 9.3 miles, depending on the latitude) (Maurer et al. 2002). In addition to the gridded data set, weather station data from the Global Historical Climatology Network were used to train the downscaling model. Weather stations were required to have at least two decades of continuous daily observations in order to robustly sample from the range of natural climate variability and to avoid overfitting model results (Hayhoe 2010b). This data set was chosen for several reasons. First, it covered the entire United States, and thus allowed a consistent data set to be used in this and other regional vulnerability assessments being conducted simultaneously. Second, it included downscaled projections for the A1FI emissions scenario, which is the scenario that most closely matches current trends in global greenhouse gas emissions (Raupach et al. 2007). Third, the availability of data at daily time steps was advantageous because it was needed 59 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG for some impact models used in this report. Fourth, the quantile regression method is more accurate at reproducing extreme values at daily time steps than simpler statistical downscaling methods (Hayhoe 2010b). Finally, the 7.5-mile grid scale resolution was fine enough to be useful for informing land management decisions. A disadvantage is that some cells within the assessment area represent highly complex landforms with steep elevation gradients from valleys to ridges, but cannot account for the local microclimates or changes in microclimates on a smaller scale. Summarized projected climate data are shown in Chapter 4. To show projected changes in temperature and precipitation, we calculated the average daily mean, minimum, and maximum temperatures for each month for three 30-year time periods (2010 through 2039, 2040 through 209, 2070 through 2099). The monthly averages were used to calculate seasonal and annual values. Mean sums of average daily precipitation were also calculated for each season and annually for the same time periods. We then subtracted these values from the corresponding baseline climate average (1971 through 2000) to determine the departure from current climate conditions. Historical climate data used for the departure analysis was taken from ClimateWizard (Girvetz et al. 2009). Chapter 3 includes more information about the observed climate data from ClimateWizard. The downscaled future climate projections were also used in each of the forest impact models described below. This consistency in future climate data allows for more effective comparison across different model results. These models generally require monthly precipitation and temperature values as inputs. They also operate on grid scales that may be larger or smaller than the grid scale of the downscaled data set, and grid scales were adjusted accordingly. iMPACT MoDELS Downscaled climate projections from GCMs provide important information about future climate, but they tell us nothing about how climate change might affect soil moisture, hydrology, forest composition, productivity, or interactions between these factors. Other models, commonly called impact models, are needed to project impacts on physical and biological processes (Fig. 13). Impact models use downscaled GCM projections as inputs, as well as information about tree species, life history traits of individual Figure 13.—Steps in the development of climate impact models using projecions from general circulaion models (GCMs) and speciic steps taken in this assessment. 0 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG species, and soil types. Several different models are used to simulate impacts on species and forest ecosystems. These models generally fall in one of two main categories: species distribution models (SDMs) and process models. In this assessment, we used one species distribution model, the Climate Change Tree Atlas (Landscape Change Research Group 2014), and two process models, LINKAGES (version 2.2; Wullschleger et al. 2003) and LANDIS PRO (Wang et al. 2013). These models operate at different spatial scales and provide different kinds of information. We chose them because they have been used to assess climate change impacts on ecosystems in our geographic area of interest, and have stood up to rigorous peer review in scientific literature. MoDELS FoR ASSESSiNG FoREST ChANGE Species distribution models establish a statistical relationship between the current distribution of a tree species and key attributes of its habitat. This relationship is used to predict how the range of the species will shift as climate change affects those attributes. Species distribution models, such as the Tree Atlas, are much less computationally expensive than process models, so they can typically provide projections for the suitable habitat of many species over a larger area. There are some caveats that users should be aware of when using them, however (Wiens et al. 2009). These models use a species’ realized niche instead of its fundamental niche. The realized niche is the actual habitat a species occupies given predation, disease, and competition with other species. A species’ fundamental niche, in contrast, is the habitat it could potentially occupy in the absence of competitors, diseases, or herbivores. Given that a species’ fundamental niche may be greater than its realized niche, SDMs may underestimate current niche size and future suitable habitat. In addition, species distributions in the future might be constrained by competition, disease, and predation in ways that do not currently occur. If so, SDMs could overestimate the amount of suitable habitat in the future. Furthermore, fragmentation or other physical barriers to migration may create obstacles for species otherwise poised to occupy new habitat. Therefore, a given species might not actually be able to enter the assessment area in the future, even if an SDM like the Tree Atlas projects it will gain suitable habitat. Additionally, the Tree Atlas does not suggest that existing trees will die if suitable habitat moves out of an area. Rather, this is an indication that they will be living farther outside their ideal range and will be exposed to more climate-related stress. In contrast to SDMs, process models, such as LANDIS PRO and LINKAGES, simulate ecosystem and tree species dynamics based on mathematical representations of physical and biological processes. Process models can simulate future change in tree species dispersal, succession, biomass, and nutrient dynamics over space and time. Because these models simulate spatial and temporal dynamics of a variety of complex processes and operate at a finer pixel size, they typically require more computational power than SDMs. Therefore, fewer species can be modeled compared to SDMs. Process models also have several assumptions and uncertainties that should be taken into consideration when applying results to management decisions. Process models rely on empirical and theoretical relationships that are specified by the modeler. Any uncertainties in these relationships can be compounded over time and space, leading to potential biases. Although useful for projecting future changes, both process models and SDMs share some important limitations. They assume that species will not adapt evolutionarily to changes in climate. This assumption may be true for species with long generation times (such as trees), but some short-lived species may be able to adapt even while climate is rapidly changing. Both types of models may also magnify the uncertainty inherent in their input data. Data on the current distribution of trees, site 1 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG characteristics, and downscaled GCM projections are estimates that add to uncertainty. No single model can include all possible variables, and there are “unknown unknowns”; thus there are important inputs that will be excluded from individual models. In this assessment, competition from understory vegetation, herbivory, and pest outbreaks are a few of the processes excluded from the impact models. Given these limitations, it is important for all model results to pass through a filter of local expertise to ensure that results match with reality on the ground. Chapter  and Appendix 5 explain the expert elicitation process for determining the vulnerability of forest ecosystems based on local expertise and model synthesis. Climate Change Tree Atlas The Climate Change Tree Atlas incorporates a diverse set of information about potential shifts in the distribution of tree species habitat in the eastern United States over the next century (Landscape Change Research Group 2014). The species distribution model DISTRIB measures relative abundance, referred to as importance value, for 134 eastern tree species. The model then projects future importance values and suitable habitat for individual tree species by using downscaled GCM data readjusted to a 12.4-mile grid of the eastern United States (east of the 100th meridian) (Landscape Change Research Group 2014). The DISTRIB model uses inputs of tree abundance, climate, and the environment to simulate species habitat. Tree abundance is estimated from the U.S. Forest Service’s Forest Inventory and Analysis (FIA) data plots (Miles et al. 200). Current and future climates are simulated from the most recent downscaled climate data created by Hayhoe and colleagues (Hayhoe 2010a) for two GCMs (GFDL and PCM) and two emissions scenarios (A1FI and B1) (see Chapter 4 for maps of downscaled climate data for the assessment area). Inputs characterizing land use, fragmentation, climate, elevation, soil class, and soil properties were obtained from various agencies and data clearinghouses to provide the 38 predictor variables (Table 11) (Iverson et al. 2008b, Riitters et al. 2002). The reliability of individual habitat models is evaluated through the calculation of a model reliability score, which is based on statistically quantified measures of fitness (methods are fully described in Matthews et al. [2011a]). Steep slopes, prone to soil erosion and slippage. Photo by Patricia Butler, Northern Insitute of Applied Climate Science (NIACS) and Michigan Tech, used with permission. 2 Each tree species is further evaluated for additional factors not accounted for in the statistical models (Matthews et al. 2011b). These modifying factors (Appendix 4) are supplementary information on life history characteristics such as dispersal ability or ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Table 11.—Parameters used to predict current and future tree species habitat (Iverson et al. 2008b) Land use and fragmentaion (%) Cropland Nonforest land Forest land Water Fragmentaion index (Riiters et al. 2002) Climate (°C, mm) Mean annual temperature Mean January temperature Mean July temperature Mean May through September temperature Annual precipitaion Mean May through September precipitaion Mean diference between July and January temperature Elevaion (m) Elevaion coeicient of variaion Minimum elevaion Maximum elevaion Average elevaion Range of elevaion fire tolerance as well as information on sensitivity to disturbances such as pests and diseases that have had negative effects on the species. This supplementary information allows us to identify when an individual species may do better or worse than model projections suggest. There are important strengths and limitations of the Tree Atlas that should be considered when interpreting results. Importantly, DISTRIB projects where the habitat suitability may potentially change for a particular species, but does not project where the species may actually occur by a certain time. The actual rate of migration into the new suitable habitat will be influenced by large time lags, dispersal and establishment limitations, and availability of refugia. The FIA data plots are nonbiased and extensive across the assessment area, but are spatially sparse at Soil class (%) Alisol Aridisol Enisol Histosol Incepisol Mollisol Spodosol Ulisol Verisol Soil property Soil bulk density (g/cm3) Potenial soil producivity (m3/ha imber) Percent clay (<0.002 mm size) Soil erodibility factor Soil permeability rate (cm/h) Percent soil passing sieve no. 10 (coarse) Soil pH Depth to bedrock (cm) Percent soil passing sieve no. 200 (ine) Soil slope (%) of a soil component Organic mater content (% by weight) Total available water capacity (cm) a 12.4-mile resolution. Species that are currently rare on the landscape are often undersampled in the FIA data, and consequently have lower model reliability. Likewise, species that are currently abundant on the landscape usually have higher model reliability. The methods assume the species are in equilibrium with the environment, and do not account for species that rapidly change distributions (e.g., invasive species). The models do not account for biological or disturbance factors (e.g., competition or fire) that affect species’ abundance. Thus, the modifying factors are provided as a supplement to the model output to help address these deficiencies. For this assessment, DISTRIB uses the GFDL A1FI and PCM B1 climate scenarios. The results provided in Chapter 5 are now available from the online Climate Change Tree Atlas (www.nrs.fs.fed.us/atlas) under “Regional Assessments.” 3 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG LiNKAGES The LINKAGES model (version 2.2; Wullschleger et al. 2003) is a forest succession and ecosystem dynamics process model modified from an earlier version of LINKAGES (Pastor and Post 1985). The LINKAGES model integrates establishment and growth of individual tree species over 30 years on a plot from bare ground, incorporating ecosystem functions such as soil-water balance, litter decomposition, nitrogen cycling, soil hydrology, and evapotranspiration. Inputs to the model include climate variables (e.g., daily temperature, precipitation, wind speed, and solar radiation), soil characteristics (e.g., soil moisture capacity and percent rock, sand, and clay for multiple soil layers), and biological traits for each tree species (e.g., growth rate and tolerance to cold and shade). A full list of model inputs is presented in Table 12. Outputs to the model include tree species composition, number of stems, biomass, leaf litter, available nitrogen, humus, and organic matter, as well as hydrologic dynamics such as runoff. Unlike the LANDIS PRO model (below), LINKAGES is not spatially dynamic, and does not simulate tree dispersal or any other spatial interaction among grid cells. Simulations are done at yearly time steps on multiple 0.2-acre circular plots, which correspond to the average gap size when a tree dies and falls over. Typically, the model is run for a specified number of plots in an area of interest, and results are averaged to determine relative species biomass and composition across the landscape over time. For this assessment, LINKAGES simulates changes in tree species establishment probability over the next century for 23 common tree species within the Table 12.—Parameters used in the LiNKAGES model Locaion Laitude, longitude Climate (daily) Total daily precipitaion Daily minimum temperature Daily maximum temperature Daily total solar radiaion Mean monthly wind speed Soil Field capacity for 12 soil layers Wiling point for 12 soil layers Hydrological coeicients for 12 soil layers (based on percent sand and clay) Organic mater (Mg/ha) Nitrogen (Mg/ha) Percent rock for 12 soil layers Tree speciesa Total annual degree day maximum and minimum (Moscow Forestry Sciences Laboratory 2014) Height and diameter growth equaion coeicients (Miles et al. 2006) Typical maximum mortality age (Loehle 1988) Frost tolerance (Moscow Forestry Sciences Laboratory 2014) Shade tolerance Drought tolerance Nitrogen equaion coeicients (Natural Resources Conservaion Service 2014b, Post and Pastor 1996) Sprout stump number and minimum and maximum diameter Mineral or organic seed bed Maximum seeding rate Crown area coeicients Root:shoot raio by species Leaf liter quality class Foliage retenion ime Leaf weight per unit crown area a 4 From Post and Pastor (1996) unless noted otherwise. ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Central Appalachians region. The model projects changes in forest composition by using downscaled daily mean temperature and precipitation under GFDL A1FI and PCM B1, and compares these projections with those under a current climate scenario (i.e., the climate during 1990 through 2009) at a future time period. One hundred and fifty-six 0.2-acre virtual plots were parameterized in LINKAGES; this number represents 1 plot for each of  landforms in 2 ecological subsections. There are important strengths and limitations of LINKAGES that should be considered when interpreting results. Section-level estimates were derived from the weighted average of landforms in a subsection and the weighted average of subsections in a section. Most of the 15 plots were located at the geographic center of a subsection, which provided climate variables that represented average conditions for that subsection. However, plot locations were modified for subsections with large elevation gradients, such as the Northern High Allegheny Mountain subsection within the Allegheny Mountains section (elevation 1,7 to 4,7 feet). For these subsections, the geographic center was often located at either the highest elevation or the lowest elevation, which skewed the temperature values to appear colder (at high elevation) or warmer (at low elevation) than the average of the subsection. Therefore, plots were located at a representative point at a mid-elevation. This approach better reflects the average climate conditions for a subsection, but nevertheless fails to fully address the elevation gradient and associated climate conditions for tree species within mountainous sections. Therefore some species that occur only at the upper or lower end of an elevation gradient may appear to have lower growth potential than expected, because the results represent the average of the entire subsection. LANDiS PRo The LANDIS PRO model (Wang et al. 2014) is a spatially dynamic process model that simulates species-, stand-, and landscape-level processes. It is derived from the LANDIS model (Mladenoff 2004), but has been modified extensively from its original version. The LANDIS PRO model can simulate very large landscapes (millions of acres) at relatively fine spatial and temporal resolutions (typically 200 to 300 feet and 1- to 10-year time steps). One new feature of the model compared to previous versions is that inputs and outputs of tree species data include tree density and volume and are compatible with FIA data. Thus, the model can be directly initialized, calibrated, and validated with FIA data. This compatibility ensures the starting simulation conditions reflect what is best known on the ground and allows the modelers to quantify the uncertainties embedded in the initial data. Species-level processes include seedling germination and establishment, growth, vegetative reproduction, and tree mortality. Species-level processes are simulated from known life history characteristics and empirical equations. Stand-level processes include competition and succession. Landscapelevel processes include fire, wind, insect outbreaks, disease, invasive species, harvesting, silviculture, and fuels treatments. The LANDIS PRO model stratifies the landscape into land types based on topography, climate, soil, and other environmental characteristics. Within a land type, species establishment and resource availability are assumed to be similar. Combined with anthropogenic and natural disturbances, these land type-specific processes are capable of simulating landscape heterogeneity, time, and space. 5 ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Basic inputs to the LANDIS PRO model include maps of species composition, land types, stands, management areas, and disturbance areas. In addition, species characteristics such as longevity, maturity, shade tolerance, average seed production, and maximum diameter at breast height are given as inputs into the model. A software program, Landscape Builder, is used to generate the species composition map (Dijak 2013). Landscape Builder uses the FIA unit map, national forest type map, national size class map, the National Land Cover Dataset (U.S. Geological Survey 2011), and landform maps to assign the number of trees by age cohort and species to each grid cell. Landform maps specify the slope, aspect, and landscape position to replicate the complex topography of the assessment area (Fig. 14). Initialized landscapes are compared to FIA data at both the landscape and land type scale. Species models are calibrated by adjusting the model input parameters until simulation results match FIA data. In this assessment, the initial landscape was simulated from 1978 through 2003 data, and the number of trees and basal area by species was compared to 2003 FIA data. Results of the model predictions are validated by comparing simulations from 1978 through 2008 to the FIA data of 2008. The Landscape Builder model was also validated to verify that the theories and assumptions in LANDIS PRO are valid. The calibrated and validated model was further validated by comparing long-term simulations (150 years) to Gingrich stocking charts and Reineke density diagrams to verify that stand development processes and relationships were adequately simulated (Wang et al. 2013). Figure 14.—Example landform map (e.g., subsecion 890) used in landscape iniializaion in the LANDIS PRO model (Dijak 2013).  ChAPTER 2: CLiMATE ChANGE SCiENCE AND MoDELiNG Basic outputs in LANDIS PRO for a species or species cohort include biomass, age, and carbon. Disturbance and harvest history can also be simulated across space and time. The spatially dynamic nature of the model and its fine spatial resolution are unique advantages of LANDIS PRO compared to LINKAGES (described above) and statistically based models such as DISTRIB. Disadvantages of LANDIS PRO are that it is too computationally intensive to be run for a large number of species (in contrast to DISTRIB) and does not account for ecosystem processes such as nitrogen cycling or decomposition (in contrast to LINKAGES). For this assessment, LANDIS PRO simulates changes in basal area and trees per acre on 8-foot grid cells over the next century for 1 dominant tree species across the Central Appalachians region. The model projects changes in forest composition by using downscaled daily mean temperature and precipitation from the GFDL A1FI and PCM B1 climate scenarios, and compares these projections with those under a current climate scenario. ChAPTER SuMMARY Temperatures have been increasing in recent decades at global and national scales, and the overwhelming majority of climate scientists attribute this change to increases in greenhouse gases from human activities. Even if dramatic changes are made to help curtail greenhouse gas emissions, these greenhouse gases will persist in our atmosphere for decades to come. Scientists can model how these increases in greenhouse gases may affect global temperature and precipitation patterns by using GCMs. These large-scale climate models can be downscaled and incorporated into other types of models that project changes in forest composition and ecosystem processes. Although there are inherent uncertainties in what the future holds, all of these types of models can help us frame a range of possible futures. This information can then be used in combination with the local expertise of researchers and managers to provide important insights about the potential effects of climate change on forest ecosystems. There are important strengths and limitations of LANDIS PRO that should be considered when interpreting results. This model assumes that historical successional dynamics are held constant into the future. It is also assumed that the resource availability by land type was able to capture the effects of landscape heterogeneity at the 8-foot resolution. Species that are currently rare on the landscape are often undersampled in the FIA data, and consequently have lower model reliability. Species that are currently abundant on the landscape have higher model reliability. 7 ChAPTER 3: oBSERVED CLiMATE ChANGE As discussed in Chapter 1, climate is one of the principal factors that has determined the composition and extent of forest ecosystems in the Central Appalachians over the past several thousand years. This chapter describes the climate trends in the assessment area that have been observed over the past century, including documented patterns of climate-related processes and extreme weather events. It also presents evidence that ecosystems in the Central Appalachians are already exhibiting signals that they are responding to shifts in temperature and precipitation. CuRRENT CLiMATE The existing climates within the Central Appalachians are strongly influenced by atmospheric weather, latitude, topography, and changes in elevation (Chapter 1). Lake-effect precipitation is a factor in the north and west, whereas dramatic changes in elevation are responsible for orographic effects on rain and snow in the mountainous regions. This heterogeneity of climates across the region can be seen at finer scales, but is often lost in broadscale averages. Temperature and precipitation at weather stations in the Central Appalachians region have been recorded for more than 100 years. The average temperature and precipitation across the assessment area was examined by using the ClimateWizard Custom Analysis tool (ClimateWizard 2013, Girvetz et al. 2009). Data for the tool are derived from PRISM (Parameter-elevation Regressions on Independent Slopes Model) (Gibson et al. 2002), which models historical measured point data onto a continuous 2.5-mile grid over the entire United States. Temperature and precipitation data were used to derive annual, seasonal, and monthly values for the 30-year average (also referred to as the “climate normal”) for 1971 through 2000 (Table 13, Figs. 15 and 1) and for each section (Appendix 2). Table 13.—Annual and seasonal mean, minimum, and maximum temperature and total precipitaion for 1971 through 2000 (ClimateWizard 2013) Season Annual Winter (Dec-Feb) Spring (Mar-May) Summer (Jun-Aug) Fall (Sep-Nov) 8 Mean temperature (°F) Minimum temperature (°F) Maximum temperature (°F) Mean precipitaion (inches) 51.1 31.2 50.2 70.1 53.0 40.0 21.7 38.0 58.5 41.6 62.3 40.7 62.4 81.6 64.4 43.1 9.2 11.5 12.7 9.7 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 15.—Thirty-year annual and seasonal averages of mean, minimum, and maximum temperatures across the assessment area from 1971 through 2000. Data source: ClimateWizard (2013). 9 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 16.—Thirty-year averages of mean annual and seasonal precipitaion across the assessment area from 1971 through 2000. Data source: ClimateWizard (2013). 70 ChAPTER 3: oBSERVED CLiMATE ChANGE hiSToRiCAL TRENDS iN TEMPERATuRE AND PRECiPiTATioN The Central Appalachians region has experienced changes in temperature and precipitation over the past 111 years, and the rate of change appears to be increasing. Long-term trends from 1901 through 2011 were examined by using the ClimateWizard Custom Analysis tool to gain a better understanding of how climate has been changing (Appendix 2). Trends in annual, seasonal, and monthly temperature (mean, minimum, and maximum) and total precipitation were examined both for the entire assessment area, and separately for each ecological section within the assessment area (Tables 14 and 15). Long-term trends show that some aspects of the climate have been changing. Accompanying tables and figures present the change over the 111-year period estimated from the slope of the linear trend. In the following text we highlight increasing or decreasing trends for which we have moderate to high confidence that they did not occur by chance. For more information regarding confidence in trends and the PRISM data, refer to Appendix 2. Observed changes in other ecological indicators are often described on a statewide basis because finer resolution data were not available, unless otherwise indicated. Temperature Between 1901 and 2011, annual mean temperatures fluctuated from year to year by several degrees. The coolest year on record was 1917, and the warmest year on record was 1921 (Fig 17). Many of the highest temperatures on record were between 1921 and the mid-1950s, and there was a cool period in the 190s and 1970s. Temperatures appear to be increasing in the past few decades, but they are not as high as were experienced in the mid-20th century. Although annual mean temperatures increased both globally and across the United States over the same time period, the increase in the assessment area was very small (0.5 °F) (Fig. 17). Seasonal mean temperatures did not change overall (Table 14), but there were several trends when monthly mean temperatures were examined (Fig. 18). April mean temperatures increased by 2.4 °F. August mean temperatures increased by 1.2 °F, and November mean temperatures increased by 2.3 °F. Although the direction of change appears negative for the maximum temperatures in all seasons except spring, trends in maximum temperatures were small enough that they may have occurred by chance. Maximum temperatures increased the most in April (3.2 °F), and decreased the most in September (-2.1 °F), October (-2 °F), and July (-1.2 °F). Annual minimum temperature increased by 1.1 °F. Minimum temperatures also increased in summer (1. °F) and fall (1.4 °F) (Fig. 18). Minimum temperatures increased the most in April (1. °F), June (1.4 °F), July (1.3 °F), August (2.1 °F), and November (2.8 °F). April and November are notable because both minimum and maximum temperatures increased in those months. 71 ChAPTER 3: oBSERVED CLiMATE ChANGE Table 14.—Change in annual and seasonal mean temperatures and precipitaion from 1901 to 2011 in the assessment area Season Annual Winter (Dec-Feb) Spring (Mar-May) Summer (Jun-Aug) Fall (Sep-Nov) a Mean temperature (°F) Minimum temperature (°F)a Maximum temperature (°F) Precipitaion (inches) 0.5 0.3 0.8 0.6 0.3 1.1 0.8 0.6 1.6 1.4 -0.1 -0.1 0.9 -0.4 -0.7 1.7 -1.0 0.7 -0.3 2.3 Values in boldface indicate less than 10-percent probability that the trend could have occurred by chance alone. Table 15.—Change in annual and seasonal mean temperatures and precipitaion from 1901 to 2011 by ecological secion within the assessment area (ClimateWizard 2013) Secion Season Mean temperature (°F)a Minimum temperature (°F) Maximum temperature (°F) Precipitaion (inches) 221E Annual Winter Spring Summer Fall 0.3 -0.1 0.4 0.4 0.3 0.7 0.4 0.1 1.2 1.0 -0.1 -0.5 0.8 -0.4 -0.5 1.3 -1.2 0.4 -0.1 2.3 221F Annual Winter Spring Summer Fall 0.5 0.9 1.1 0.2 -0.1 1.0 1.3 0.9 1.1 0.7 0.1 0.4 1.3 -0.7 -0.8 4.2 0.1 0.8 1.0 2.4 M221A Annual Winter Spring Summer Fall 1.4 1.7 1.8 1.6 0.6 2.1 2.2 1.6 2.6 2.1 0.7 1.1 1.9 0.6 -0.9 2.0 -0.4 1.7 -2.2 2.8 M221B Annual Winter Spring Summer Fall 0.8 0.4 1.1 1.0 0.6 1.9 0.8 1.6 2.6 2.4 -0.3 0.0 0.6 -0.7 -1.2 0.0 -1.8 0.9 -1.3 2.2 M221C Annual Winter Spring Summer Fall 0.3 -0.4 0.3 0.7 0.6 1.1 -0.1 0.4 1.8 2.1 -0.5 -0.6 0.1 -0.5 -0.9 0.8 -1.6 0.7 -0.4 2.1 a Values in boldface indicate less than 10-percent probability that the trend could have occurred by chance alone. 72 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 17.—Annual mean temperature across the assessment area from 1901 through 2011. The blue line represents the rolling 5-year mean. The red regression line shows the trend across the enire ime period (0.005 °F per year; p = 0.28). Data source: ClimateWizard (2013). 73 ChAPTER 3: oBSERVED CLiMATE ChANGE 4.0 * 3.0 * * * * 2.0 * * T em p eratu re c h ang e (°F ) * * * 1.0 M ean 0.0 M inim um M ax im um -1.0 * -2.0 * * -3.0 -4.0 Ja n u a ry Fe b ru a ry M a rch A p ril May Ju n e Ju ly A u g u st S e p te m b e r O cto b e r N o ve m b e r D e ce m b e r M o nth Figure 18.—Change in monthly mean, minimum, and maximum temperatures across the assessment area from 1901 through 2011. Asterisks indicate there is less than 10-percent probability that the trend could have occurred by chance alone. Data source: ClimateWizard (2013). 74 ChAPTER 3: oBSERVED CLiMATE ChANGE Temperature trends also differed geographically across the assessment area, with some areas warming or cooling more than others (Table 14). In general, the easternmost sections (M221A and M221B) have changed more than other sections (see Chapter 1 for a map showing ecological sections). Annual mean temperatures increased by 1.4 °F in M221A and by 0.8 °F in M221B. There were no trends in mean winter and fall temperatures in any section of the assessment area. Mean spring temperatures increased in 221F, M221A, and M221B. Mean summer temperatures increased in M221A, M221B, and M221C. Minimum temperatures increased the most in M221A and M221B. In M221A, minimum temperatures increased annually and in all seasons. In M221B, minimum temperatures increased annually and in all seasons except winter. In both M221A and M221B, minimum temperatures increased the least in spring (1. °F) and the most in summer (2. °F). Maximum temperatures increased in two sections, both in spring (221F and M221A). Maximum temperature cooled significantly only in fall and only in Section M221B. Spatially interpolated trends in temperature are available through 2011 and are presented in Fig. 19. Stippling on the maps indicates trends which have moderate to high probability that they did not occur by chance. Spatial analysis showed that increases in annual temperatures ranged from 1 to 4 °F over large portions of the assessment area, whereas decreases were observed in only a few isolated locations, indicated by the stippling. The greatest increases are observed in minimum temperatures, with increases of up to  °F consistently appearing in Sections M221A and M221B and increases of up to 3 °F appearing in widespread areas throughout the assessment area. The greatest decreases are observed in maximum temperatures, with widespread areas cooling by as much as 5 °F. These observed decreases in summer maximums may be evidence of a regional “warming hole” (see section on “Regional Patterns Contributing to Local Trends”). Precipitaion From 1901 through 2011, mean annual precipitation within the assessment area fluctuated by as much as 20 inches from year to year (ClimateWizard 2013) (Fig. 20). The driest year on record for the assessment area as a whole was 1929. Precipitation was lower than the long-term average during distinct periods over the last century, including a few years during the “Dust Bowl” era of the 1930s, a dry spell from 190 through 199, and 1987. The four wettest years on record occurred during the past 20 years. The time series of annual precipitation for the assessment area displays high variability from year to year and the surge in precipitation at the beginning of the 21st century may be driving an upward trend, although there are not enough data years from the 21st century to determine whether the trend is real or due to chance (Box 7). Because of the large interannual variability of precipitation averaged across the assessment area, any positive or negative trend observed in mean annual or seasonal precipitation in the assessment area was small enough that it could have occurred by chance, except for fall (Table 14). Fall precipitation increased by 2.3 inches from 1901 to 2011. Several trends were observed in monthly mean precipitation (Fig. 22). When averaged across the entire assessment area, precipitation increased in May (0.9 inches), September (0.9 inches), and November (1.2 inches). 75 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 19.—Annual and seasonal change in mean, minimum, and maximum temperatures across the assessment area from 1901 through 2011. Sippling indicates there is less than 10-percent probability that the trend could have occurred by chance alone. Data source: ClimateWizard (2013). 7 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 20.—Annual mean precipitaion across the assessment area from 1901 through 2011 (ClimateWizard 2013). The blue line represents the rolling 5-year mean. The red regression line shows the trend across the enire period (0.015 inches per year; p = 0.26). Data source: ClimateWizard (2013). 77 ChAPTER 3: oBSERVED CLiMATE ChANGE Box 7: Climate Changes over the 21st Century In this chapter, we present changes in climate over the enire historical record for which spaially interpolated data trends are available for the assessment area. Looking across the enire record is helpful in detecing long-term changes, but it can also obscure short-term trends. In fact, the long-term trend is made up of shorter periods of warming and cooling, depending on the ime period analyzed. The period from 2001 to 2012 was the warmest on record for the world, North America, and the United States (Blunden and Arndt 2012, World Meteorological Organizaion 2011). Statewide averages for the early 21st century can be explored within the enire climate record (1895 to 2012) through the Naional Climaic Data Center’s Climate at a Glance maps (Naional Oceanic and Atmospheric Administraion [NOAA] 2014b). Temperatures across Ohio were above the long-term average for 7 years during this period, and below average for 4 years (NOAA 2014d). Ohio experienced its record warmest temperatures since the 1920s in 1998 and again in 2012 (NOAA 2014d) (Fig. 21). Maryland experienced its second warmest year in 2012 (its warmest year was in 1998). Since 2000, only 1 year was below the long-term average in Maryland, and the rest were above average. West Virginia experienced its record low for average annual temperature in 1917, followed by its record high in 1921. West Virginia displays an enormous amount of variaion from year to year, with the third warmest temperature in 2012. Precipitaion has also changed dramaically during 2001 through 2012. Ohio experienced its wetest year since 1895 in 2011. Since 2000, 7 years have been above average or much above average, whereas only 4 years were slightly below average (NOAA 2014d). Maryland and West Virginia also follow this patern; both experienced their wetest year in 2003, and both have had more above-average years than below average. And what about the “warming hole” paterns of low summer temperatures and high spring and summer precipitaion? Across the assessment area, summer temperatures during the 21st century were much higher than the long-term average for the area, with record warming in Maryland (NOAA 2014d). Although it is too early to determine a trend, the recent warming temperatures suggest a possible reversal of the “warming hole.” Summer precipitaion trends have not changed markedly in the assessment area over the 21st century, but spring precipitaion has been higher than average (NOAA 2014d). Overall, the climate informaion from the 21st century seems to be consistent with the trends over the past century in some ways but not others. The area is geing generally weter, and the 1930s coninues to be the warmest decade on record. Figure 21.—Annual mean temperatures for Ohio from 1895 through 2011. Image courtesy of NOAA (2014d). 78 ChAPTER 3: oBSERVED CLiMATE ChANGE 1.5 * C h an ge in p recipitation (inches) 1.0 * * 0.5 0.0 -0.5 -1.0 J an uary F ebruary M arc h Ap ril M ay J un e J uly Aug us t Sep tem ber O c to ber N o v em ber D ec em ber M o n th Figure 22.—Change in monthly mean precipitaion within the assessment area from 1901 through 2011. Asterisks indicate there is less than 10-percent probability that the trend could have occurred by chance alone. Data source: ClimateWizard (2013). When we examined changes at the ecological section level, trends emerged in some areas (Table 14). In general, the greatest increase in precipitation was observed in Ohio (Section 221F), where an increase of 2.4 inches was observed in fall, contributing to a total increase of 4.2 inches in annual precipitation. Precipitation increased in fall in every section, with the greatest increase in M221A (2.8 inches). In the easternmost section (M221A), summer precipitation decreased by 2.2 inches from 1901 to 2011. In southern West Virginia (M221C), winter precipitation decreased by 1. inches. Spatially interpolated trends in precipitation are available through 2011 and are presented in Figure 23. Spatial analysis showed that increases in annual precipitation ranged from 1 to 4 inches over large portions of the assessment area, whereas decreases were observed in only a few isolated locations. Precipitation has increased the most during fall, and has decreased the most during winter. Precipitation has increased over large areas in spring, but has decreased during the summer in the easternmost sections. 79 ChAPTER 3: oBSERVED CLiMATE ChANGE Figure 23.—Annual and seasonal changes in mean precipitaion from 1901 through 2011 in the assessment area. Sippling indicates there is less than 10-percent probability that the trend has occurred by chance. Data source: ClimateWizard (2013). 80 ChAPTER 3: oBSERVED CLiMATE ChANGE Regional Paterns Contribuing to Local Trends Some studies have observed a decrease in temperature, especially summer highs, in the southeastern and central United States since the 1950s, a phenomenon that has been referred to as a “warming hole” (Kunkel et al. 2013b, Meehl et al. 2012, Pan et al. 2004, Portmann et al. 2009). A recent study examined mean temperatures across the United States from 1950 through 1999 and found that decadal variations in sea-surface temperature were the most important contributor to the observed warming hole (Meehl et al. 2012). These findings are consistent with other studies that found that decreases in summer high temperature are correlated with increases in sea-surface temperatures (Kunkel et al. 200), precipitation (Pan et al. 2004, Portmann et al. 2009), aerosols (Leibensperger et al. 2012), and increased soil moisture availability (Pan et al. 2004). Further research is needed to understand the “warming hole” and its implications for the region as global air and sea surface temperatures continue to rise. An analysis of recent climate trends in the United States suggests that the warming hole may have already disappeared as the mean annual temperature has increased significantly in all states since 1970 (Tebaldi et al. 2012). Observed temperature and precipitation trends in the assessment area are consistent with the “warming hole” pattern in the regional climate. When averaged across the assessment area, maximum temperatures decreased and precipitation increased in July (Fig. 18) (ClimateWizard 2013). hiSToRiCAL TRENDS iN EXTREMES Although it can be very instructive to examine longterm trends in mean temperature and precipitation, in many circumstances extreme events can have a greater impact on forest ecosystems and the human communities that depend on them. Weather or climate extremes are defined as individual weather events or long-term patterns that are unusual in their occurrence or have destructive potential (Bader et al. 2008). These events can trigger catastrophic disturbances in forest ecosystems, along with significant socioeconomic disasters. In addition, the distribution of individual species or forest types is often controlled by particular climatic extremes. Scientists agree that climate change has increased the probability of several kinds of extreme weather events, although it is difficult to directly attribute one particular event to climate change (Coumou and Rahmstorf 2012). As mean summer and winter temperatures have increased at a national scale, the chance of experiencing unusually warm or cool seasons has become higher over the last 30 years (Hansen et al. 2012). Extreme events are difficult to analyze with standard statistical methods, so long-term studies of weather and climate trends are necessary. Extreme Temperatures Extreme temperatures can influence forest ecosystems in a variety of ways: some tree species are limited by hot growing-season temperatures, and others are limited by cold winter temperatures. Extreme temperatures may also be associated with disturbance events like drought, wildfire, ice storms, and flooding. Warmer mean temperatures are often correlated with higher extreme temperatures (Kling et al. 2003, Kunkel et al. 2008). Long-term records indicate that the number of hot days (exceeding the 95th percentile of warm temperatures) has increased across most of the contiguous United States since the 190s, a time series which excludes the hot, droughty years of the 1930s and 1950s (DeGaetano and Allen 2002; Kunkel et al. 2013b, 2013c; Perera et al. 2012; Peterson et al. 2008). The number of extreme cold days (not exceeding the 5th percentile of cold maximum and minimum temperatures) has decreased across the United States since the 190s (DeGaetano and Allen 2002). Winter maximum 81 ChAPTER 3: oBSERVED CLiMATE ChANGE and minimum temperatures across the country have increased by an average 3.5 °F over the second half of the 20th century (Peterson et al. 2008). These trends correspond to global patterns of increasing occurrence of extreme hot weather and decreasing occurrence of extreme cool weather (Hansen et al. 2012). The frequency of extreme temperatures within the assessment area is a function of latitude and elevation, with northerly and high-elevation areas likely to experience fewer hot days over 90 °F and more cold days below 0 °F than southern and low-elevation areas (Polsky et al. 2000). Intense Precipitaion Precipitation has increased over the United States by an average of 5 percent during the second half of the 20th century (Karl et al. 2009, NOAA 2014b). The assessment area is located in one of the wetter regions of the country, and some areas of the assessment area have been experiencing increases in precipitation (e.g., the Ohio portion). Similar studies corroborate precipitation increases up to 25 percent in the same area (Karl et al. 2009). From 1948 through 2011, the amount of precipitation falling during a state’s largest annual storm increased by 15 percent in Ohio, 14 percent in Maryland, and  percent in West Virginia (Madsen and Willcox 2012). The timing of precipitation events has shifted, however, and intense precipitation events have become more frequent while light rain events have not changed (Groisman et al. 2012, Kunkel et al. 2008). Throughout the Midwest during the last 40 years (including Ohio and West Virginia), there was a 50-percent increase in the frequency of days with more than 4 inches of rainfall and a 40-percent increase in the frequency of days with more than  inches of rainfall (Groisman et al. 2004, 2005, 2012). A study of the eastern United States found that heavy precipitation events are occurring more frequently in Ohio and West Virginia; heavy precipitation events that used to occur every 82 12 months are now occurring every 8.9 months (Madsen and Willcox 2012). A study of the Ohio River Basin (which includes all of the assessment area) also observed an increase in heavy rainfall from 1908 to 2007; the greatest increase was found for 1-year events (25 percent), with smaller increases for events having longer average recurrence intervals (3 percent) (Bonnin et al. 2011). A study of trends in return intervals from 1950 to 2007 also found that threshold precipitation events are occurring more frequently across the Midwest and Northeast, suggesting that extreme rainfall events are becoming more frequent, even where there have been no observed increases in total precipitation (DeGaetano and Allen 2002). Severe Thunderstorms, Tornadoes, and hurricanes Storm movement across the Central Appalachians region is generally from west to east, but strong storms from the eastern seaboard can also influence weather within the assessment area. The higher Allegheny Mountains buffer the West Virginia and Ohio portions of the assessment area, but Maryland can be heavily influenced by these east coast storms. Strong thunderstorms occur most frequently from May to August within the assessment area, and there is a general increase in frequency and expansion northward and eastward as the season progresses (Robinson et al. 2013). Thunderstorm frequency is higher west of the Appalachian Mountains than in the rest of the assessment area in April and May. Based on long-term data from 189 to 1995, the assessment area averaged 35 to 45 thunderstorm days per year (Changnon 2003). A study of severe thunderstorm observations over the eastern United States identified an increase in thunderstorm frequency over the last 0 years, but it is difficult to determine whether those increases are biased by increased accuracy in storm reporting (Robinson et al. 2013). ChAPTER 3: oBSERVED CLiMATE ChANGE Tornadoes also affect the assessment area, although less frequently than thunderstorms. Most of these tornadoes occur within Ohio (17 tornadoes per year on average), with occasional occurrences in West Virginia (2) and Maryland (7) (National Weather Service 2012). Although the number of tornadoes observed in a year appears to be increasing, the slightly positive trend is biased due to increased technology and reporting success, such as the introduction of Doppler radar technology in the 1990s (NOAA 2013a). The increase in tornado occurrence is observed in only the weakest tornadoes, and there is no evidence of increasing frequency of stronger tornadoes (Kunkel et al. 2013a). Hail is often produced during tornado weather and is more prevalent in the mountainous panhandle of Maryland due to orographic lifting (as moist air is forced into high elevations) and cooler ground temperatures, which allow for less melt on descent (Mogil and Seaman 2009). Hurricanes tracking up the Atlantic seaboard also affect the assessment area. From 1985 to 2009, four major hurricanes and more than a dozen tropical storms tracked up the eastern seaboard. As a result, much of the assessment area has been subjected to intense rain, hail, wind, and flooding, although the Allegheny Mountains buffer the Ohio portion from much of the impact (Kunkel et al. 2013c). Not every hurricane formed in the Atlantic makes landfall or affects the assessment area, but there is some evidence that the strength and frequency of hurricanes have been increasing since 1970, and that this increase is associated with warming sea surface temperatures (Holland and Webster 2007, Kunkel et al. 2008). Based on the average number of hurricanes from 1981 to 2010, the 2011 hurricane season was above average, and was the 12th above-average season since 1995 (Blunden and Arndt 2012). There is no evidence of change in the frequency of hurricanes that make landfall (Holland and Webster 2007, Kunkel et al. 2008). Hurricane Isabel made landfall in 2003, followed by Irene in 2008, and Sandy in 2013. Trends in severe weather frequency are difficult to attribute to changes in climate only; recent advances in technology, population density, and social media have contributed to increases in storm and tornado reporting (Robinson et al. 2013). Losses from catastrophic storms, defined as a storm producing more than $1 million in damages, have been used to explore trends in storm frequency and severity across the central and northeastern United States (Changnon 2011a, 2011b). Windstorms In warm months of the year, the assessment area occasionally experiences very powerful straightline windstorms, otherwise known as derechos. These events can result in substantial wind-throw disturbances in forest ecosystems. A recent example was the April 2011 storm that passed through Ohio and the northeastern border of West Virginia on its southwest-to-northeast track through the central United States. This single storm produced wind gusts of 58 to 74 miles per hour, hail, and tornadoes (NOAA 2013c). A much larger storm caused 22 deaths and widespread damage across the eastern United States, including the assessment area, on June 29, 2012 (NOAA 2013b). The average annual frequency of derechos within the assessment area decreases from Ohio (11) to Maryland (9) (Coniglio and Stensrud 2004). There is not enough evidence currently available to examine trends in derecho frequency and distribution due to limited data in the first half of the 20th century (Peterson 2000). 83 ChAPTER 3: oBSERVED CLiMATE ChANGE PhYSiCAL PRoCESSES Climate and weather patterns also drive many physical processes important for forest ecosystems. Climate-driven factors such as snowpack and soil frost can regulate annual phenology, nutrient cycling, and other ecosystem dynamics. Changes to climate-driven physical processes can result in impacts and stress on forest ecosystems that might not be anticipated from mean climate values alone. This section presents a few key trends that have been observed in the Central Appalachians and throughout the broader region. Flooding and Streamlow Although floods also depend on soil saturation, soil temperature, and drainage capabilities, floods are primarily attributed to spring snowmelt, heavy rainfall, tropical storms, and hurricanes. Floods can develop slowly as the water table rises, or quickly if large amounts of rainfall rapidly exceed moisture thresholds. Although snowpack in the Central Appalachians is generally short-lived, melting can contribute substantial volume to winter and spring peak flow and flooding (Eisenbies et al. 2007, Kochenderfer et al. 2007). Areas with steep and narrow terrain are more prone to flash flooding of the smaller rivers, streams, and tributaries (Eisenbies et al. 2007). Long-term data on flooding are difficult to interpret because of the variety of measures used to describe floods. Many floods originate from small, unmonitored watersheds, and thus go unreported (Wiley and Atkins 2010). Major regional floods can be observed through stream gauge measurements and are reported for the three states within the assessment area. Sixteen major floods have been recorded in West Virginia since 1844 (Wiley and Atkins 2010). In Maryland, 57 floods were recorded from 180 to 2004, with at least 13 of them attributed to hurricanes (Joyce and Scott A small stream meandering through hemlock forest. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 84 ChAPTER 3: oBSERVED CLiMATE ChANGE 2005). In Ohio, 38 major floods were recorded from 181 to 1990, 315 minor flood events from 2000 to 2007, and a major flood in 2011 (Ohio Emergency Management Agency 2011, Robertson et al. 2011). Damage from floods has been increasing in the Midwest in recent decades (Villarini et al. 2011). A nationwide study of streamflow between 1944 and 1993 demonstrated that baseflow and median (average) streamflow have increased at many streams in the Midwest and Mid-Atlantic (Lins and Slack 1999). More recent studies have confirmed increased annual and low streamflow from 191 to 1990, at least partially due to increased storm frequency (Groisman et al. 2004). At the same time, maximum flow (including floods) did not change (Lins and Slack 1999). Several factors complicate the explanation of trends in flood frequency. Changes in flooding frequency are driven not only by increased precipitation but also by changes in land cover and land use (Groisman et al. 2004, Jones et al. 2012, Wang and Cai 2010). In particular, human-caused land-use change over the past century has had a considerable influence on flooding frequency (Villarini et al. 2011). After these factors have been taken into account, however, studied watersheds in the Midwest still exhibited increased discharge over the past several decades, which may be attributed to climate change (Tomer and Schilling 2009). Snow and Winter Storms Cold and snowy winters are characteristic of the Central Appalachians region, which lies between two major storm tracks of the eastern United States (Hartley 1999). The assessment area experiences more snowstorms than nearby southern states, but fewer than nearby northern and eastern states (Changnon and Changnon 2007). Snowfall in the Central Appalachians is influenced by many factors including winter temperature, lake-effect weather, and elevation. Although precipitation has been increasing, the proportion that falls as snow has been decreasing (Kunkel et al. 2009a, 2009b). The ratio of snow to precipitation is strongly correlated with mean daily temperature across the United States (Feng and Hu 2007). As daily temperature increased from 1949 to 2005, the proportion of precipitation falling as snow decreased over non-lake effect areas of Ohio and most of West Virginia (Feng and Hu 2007). Decreasing trends in seasonal snowfall were also observed in the central and southern Appalachians from 193 to 1993 (Hartley 1999). Regional trends indicate that although snowfall is quite variable from year to year, the most recent 30 years have had fewer heavy snowfalls, but moreintense snowfalls when they do occur (Feng and Hu 2007). Snowfall has increased over the same period in the lake-effect area of Ohio, and in the Northern Ridge and Valley section of West Virginia and Maryland (Feng and Hu 2007). Long-term records from across the Great Lakes indicate that lake-effect snow increased gradually during the 20th century, likely due to the warming of these water bodies and the decreasing trend in lake-ice cover. Across the Midwest and Northeast, long-term records have shown that ice on inland lakes is breaking up earlier in the spring and forming later in the fall (Benson et al. 2012). Annual ice cover on Lake Erie has declined by half from 1973 to 2010 (Wang et al. 2012).The combined effect of these trends is a longer ice-free period for lakes across the region and the assessment area, including Lake Erie, which influences climate and weather in the assessment area. Drought Droughts are among the greatest stressors on forest ecosystems, and can often lead to secondary effects of insect and disease outbreaks on stressed trees and increased fire risk (Maherali et al. 200). Because droughts often affect large regions, data are available at regional and statewide scales, but not at finer scales. There is no evidence for increased drought severity, frequency, or extent on 85 ChAPTER 3: oBSERVED CLiMATE ChANGE Fall colors on the Hocking State Forest, Ohio. Photo by the Ohio Department of Natural Resources, used with permission. average across the assessment area. The Palmer Drought Severity Index (PDSI) is a soil moisture index which measures meteorological drought by calculating the cumulative departure (from the long-term mean) in moisture supply and demand (Dai et al. 2004). The Palmer Hydrological Drought Index (PHDI) measures hydrological drought based on precipitation and evaporation. Both indicators can be important in understanding the effects on groundwater supply. In North America and the United States, there has been a trend toward wetter conditions since 1950, and there is no detectable trend for increased drought based on the PDSI (Dai et al. 2004, Karl et al. 199). Another study of hydrologic trends in the United States over the last century (1915 through 2003) also observed reduced duration and severity of droughts across the Central Appalachians region as a result of increased precipitation (Andreadis and Lettenmaier 200). Statewide data (NOAA 2014d) were also explored to examine changes in the yearly and seasonal Palmer drought indices. A positive (wetter) trend from 1895 to 2013 was observed in West Virginia and Ohio annually and for each season according to both 8 indices. In Maryland, the PDSI shows no increasing or decreasing trend in annual or spring droughts, but shows that fall and winter have been getting wetter and summer has been getting drier (NOAA 2014d). The PHDI shows that winter and spring have been getting wetter, whereas annual, summer, and fall conditions have been getting slightly drier in Maryland (NOAA 2014d). Growing Season Length Growing season length is often estimated as the period between the last spring freeze and first autumn freeze (climatological growing season), but can also be estimated through the study of plant phenology (biological growing season) (Linderholm 200). A large body of research indicates that the growing season has lengthened by 10 to 20 days at global, hemispheric, and national scales, primarily due to an earlier onset of spring (Christidis et al. 2007, Easterling 2002, Linderholm 200, Parmesan 2007, Parmesan and Yohe 2003, Root et al. 2003, Schwartz et al. 200b, Zhang et al. 2007). There is evidence, however, of both positive and negative ChAPTER 3: oBSERVED CLiMATE ChANGE regional trends being dissolved into these broadscale averages. Several studies suggest that the growing season is lengthening within the assessment area, but primarily due to a later onset of fall. In fact, a recent study exploring past trends in spring onset dates in the Southeast, including the assessment area, showed that spring has been occurring later by 4 to 8 days since the 1950s (Schwartz et al. 2013). This phenomenon has been linked to the warming hole, and specifically, to processes that promote cooling during the winter (Meehl et al. 2012). Another study of the Southeast and New England also found an anomalous trend toward delayed onset of spring in nearby Virginia (Fitzjarrald et al. 2001). The onset of fall is also highly influenced by local temperature changes rather than global mean temperatures (Badeck et al. 2004). Remote sensing of vegetation patterns is one method commonly used to estimate the start, end, and length of the growing season. Studies using remote sensing have found no significant trend in the start of season, but did find that the end of season occurred later, and the total growing season lengthened by approximately 9 days from 1981 through 2008 (Jeong et al. 2011, Julien and Sobrino 2009). Another study using colddegree days and satellite imagery found a correlation between increasing midsummer temperatures and later fall senescence, which causes autumn colors (Dragoni and Rahman 2012). The authors also found that end-of-season dates varied by latitude and elevation, with earlier senescence occurring in forests at higher latitudes and elevations (Dragoni and Rahman 2012). For example, despite regional trends toward later senescence from 1989 through 2008, the end of season occurs earlier in the Appalachian range than surrounding areas (Dragoni and Rahman 2012). Increases in the growing season length are causing some noticeable changes in the timing of biological activities, such as bird migration (Box 8). Box 8. Phenological indicators of Change Changes in growing season length can be observed through studies of phenology. Phenology is the iming of recurring plant and animal life-cycle stages, such as leaf-out and senescence, lowering, maturaion of agricultural plants, insect emergence, and bird migraion. A few studies examining phenology in the Central Appalachians indicate recent changes: • In a survey of 270 lowering plants in southwestern Ohio, 60 percent showed earlier spring lowering over the period from 1976 to 2003 of about 10 to 32 days (McEwan et al. 2011). The diferences among species may be atributed to diferences in sensiivity to climate as a cue to begin lowering as opposed to other indicators such as day length. • Ten species of naive bees in the Northeast (including the enire assessment area) have been emerging an average of 10 days earlier over the last 130 years, with much of the change linked to warming trends since 1970. Bee-pollinated plants are also blooming earlier, suggesing that these generalist species are keeping pace with changes in plant phenological shits (Bartomeus et al. 2011). • The purple marin, a long-distance migratory songbird that overwinters in the assessment area, has been declining across North America and Canada (Nebel et al. 2010). Populaion declines are linked to an increasing mismatch between spring arrival date and iming of food availability (Fraser et al. 2013). A recent study tracking spring migraion from the Amazon basin to two breeding sites in Pennsylvania and Virginia found that purple marins were unable to depart earlier, migrate faster, or claim breeding sites earlier in response to earlier green-up and insect emergence. 87 ChAPTER 3: oBSERVED CLiMATE ChANGE ChAPTER SuMMARY Notable shifts have been observed in climate, extreme weather events, and phenology within the assessment area. Broad regional trends have translated into high spatial variability across the region. Mean and minimum annual, spring, and summer temperatures have increased more in the mountainous parts than in other parts of the assessment area. Minimum temperatures have generally increased, and maximum temperatures have generally decreased in parts of the assessment area. Precipitation increases were detected in the fall season in every part of the assessment area, and changes during other seasons differed with location. Summer precipitation decreased in the far eastern part of the assessment area, but remained relatively stable elsewhere. Drought indices indicate that the frequency and severity of droughts have not changed. Heavy precipitation events have become more frequent and intense. Characteristic winter conditions such as snowfall and lake ice have been diminishing with warmer temperatures. In addition, the growing season has lengthened due to later onset of fall. These trends are generally consistent with regional, national, and global observations related to anthropogenic climate change, but with subtle local differences. Ecological indicators are beginning to reflect these changes as well, as evidenced by changing arrival of migratory birds and changing phenology. Sources of information on historical climate trends and ecological indicators are listed in Box 9. Box 9. More Historical Climate Informaion Much more informaion on historical climate trends and ecological indicators for the Central Appalachians region exists than was possible to present in this chapter. Interested readers will be able to ind more informaion from the following resources: New Jersey, New York, Pennsylvania, Rhode Island, Vermont, and West Virginia. It provides highquality climate data, derived informaion, and data summaries for the Midwest. Naional Informaion State-level Informaion The Naional Climaic Data Center (NCDC) is the world’s largest acive archive of weather data. The NCDC’s Climate Data Online provides free, downloadable data from the Global Historical Climatology Network. Please note that Web addresses are current as of the publicaion date of this assessment but are subject to change. www.ncdc.noaa.gov/oa/ncdc.html Regional Informaion The Northeast Regional Climate Center is a cooperaive program between the Naional Climaic Data Center (above) and the state climate oices serving the 12-state region of Connecicut, Delaware, Massachusets, Maryland, Maine, New Hampshire, 88 www.stateclimate.org/regional.php?region= northeast State climatologists provide informaion about current and historical trends in climate throughout their states. Visit your state climatologist’s Web site for more informaion about trends and climate paterns in your paricular state: Oice of the State Climatologist for Ohio htp://www.geography.osu.edu/faculty/rogers/ statclim.html West Virginia State Climate Oice & Meteorology htp://www.marshall.edu/met/ Maryland State Climatologist Oice htp://metosrv2.umd.edu/~climate/ ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES In Chapter 3, we examined how climate has changed in the Central Appalachians region over the past 111 years, based on measurements. This chapter examines how climate may change through the end of this century, including changes in extreme weather events and other climate-related processes. General circulation models (GCMs) are used to project future change at coarse spatial scales and are then downscaled in order to be relevant at scales where land management decisions are made. In some cases, these downscaled data are then incorporated into forest species distribution models and process models (see Chapters 2 and 5). Chapter 2 more fully describes the models, data sources, and methods used to generate these downscaled projections, as well as the inherent uncertainty in making long-term projections. In this chapter, we focus on two climate scenarios for the assessment area, chosen to bracket a range of plausible climate futures. Information related to future weather extremes and other impacts is drawn from published research. mean, minimum, and maximum temperatures and total daily precipitation were downscaled to an approximately 7.5-mile grid across the United States. For all climate projections, two climate scenarios are reported: GFDL A1FI and PCM B1 (see Chapter 2). The GFDL A1FI climate scenario projects greater changes in future temperature and precipitation than the PCM B1 climate scenario (hereafter referred to simply as PCM B1 and GFDL A1FI). Although both climate scenarios are possible, GFDL A1FI matches current trends in emissions and temperature more closely than PCM B1 (Raupach et al. 2007). It is possible that the future will be different from any of the developed scenarios, and therefore it is important to consider the range of possible climate conditions over the coming decades rather than one particular scenario. The 1971 through 2000 climate averages from ClimateWizard (Girvetz et al. 2009) were used as the baseline from which future departure from current climate conditions was calculated (see Chapter 3 and Appendix 2). PRoJECTED TRENDS iN TEMPERATuRE AND PRECiPiTATioN Climate projections are presented in two ways in this chapter. In general assessment area-wide trends are described first, followed by maps that show spatial variation in these trends. When the assessment area is averaged as a whole, the projections of temperature are positive, whereas projections of precipitation are positive and negative, depending on the season and model. When climate data were averaged for each grid cell within the assessment area, groups of pixels on a map begin to show subregional climate trends, such as warming in one area and cooling in another (mainly the Allegheny Mountains section; see also Box 10). The assessment area has experienced changes in temperature and precipitation over the past 100 years, and those changes are projected to increase in intensity over the next 100 years. Projected changes in temperature and precipitation within the assessment area were examined by using a statistically downscaled climate data set for three 30-year time periods through the end of this century (2010 through 2039, 2040 through 209, and 2070 through 2099) (Stoner et al. 2012). Daily 89 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Box 10. Climate Modeling in Areas of Complex Topography Areas of complex topography, such as the Allegheny Mountains and Northern Ridge and Valley secions of West Virginia and Maryland, contain some of the highest biological diversity in the world (Hoekstra et al. 2010). Paterns of ridges, valleys, slope, rainshadow efects, cold air pooling, and other ine-scale processes create a complex suite of ecological niches with various temperature and moisture regimes which may actually provide the assessment area with addiional resilience to changes (Anderson and Ferree 2010). Terrain creates various levels of decoupling between the climate experienced at a site and the broad climate trends for any given region (Dobrowski 2011, Fridley 2009). Precipitaion paterns in mountainous areas are paricularly diicult to model, owing to the complexity of atmospheric circulaion, wind speed, rainshadow efects, and orographic liting of moisture to higher elevaions. Although we can use the downscaled climate data at the regional level to gain an understanding of broad-scale trends, staisical downscaling oten does not capture landscape heterogeneity seen in some porions of the assessment area. Although few studies have invesigated iner scale modeling of mountain ranges in the United States, there have been some studies that may shed light on how downscaled climate models may be overesimaing or underesimaing temperature and precipitaion trends at various elevaions and landscape posiions. A study in the Oregon Cascades, which is prone to cold-air pooling similar to the Allegheny Mountains, found that temperatures Temperature The assessment area is projected to experience substantial warming over the 21st century, especially for GFDL A1FI (Fig. 24). Early-century (2010 through 2039) temperature increases are projected to be relatively small when averaged across the 90 in sheltered valley botoms are decoupled from the free atmosphere, and consequently are somewhat bufered from changes projected for the whole study area (Daly et al. 2010). Modeled warming of 4.5 °F at closely spaced sites simulated temperature diferences of up to 10.8 °F between low-elevaion valleys and high-elevaion ridge tops. In a study of mountainous terrain at the Hubbard Brook Experimental Forest in New Hampshire, three climate models overesimated observed precipitaion by 20 percent for the period 1979 through 2008 (Campbell et al. 2010), so that future projected values were corrected downward by 20 percent. A study in the southern Appalachians found that the winter northwest low-level air low is nearly perpendicular to the southwest-northeast mountain range, producing orographic liting and subsequent snowfall on northwest slopes and higher elevaions, despite warmer temperatures at lower elevaions (Perry and Konrad 2006). These studies suggest that there are diiculies in accurately modeling areas with complex topography and rapid elevaion change. Regional climate models have not performed as well as in areas of relaive homogeneity, and some correcion may be necessary to account for elevaion, slope, aspect, and relaive exposure or isolaion from the elements. Finerresoluion modeling would help idenify biases in the data based on these factors. Unil such ine modeling eforts can be executed, the coarse-resoluion data sets used in this assessment can provide a broad foundaion of plausible future climates from which to consider the caveats above. assessment area, with little change projected for PCM B1 (0.8 °F) and a modest increase of 2 °F for GFDL A1FI (Fig. 24, Table 1). Projections of temperature do not diverge substantially for the two future scenarios until mid-century (2040 through 209), when much larger temperature increases are ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES projected for GFDL A1FI than PCM B1 through the end of the century. Compared to the 1971 through 2000 baseline period, the average annual temperature at the end of the century is projected to increase by 1.9 °F for PCM B1 and by 7.8 °F for GFDL A1FI (Table 1). Seasonal changes follow this pattern, with modest changes projected during the early century, and the highest temperature increases projected for GFDL A1FI at the end of the century (see Appendix 3 for projected changes in mean, minimum, and maximum temperatures during the early, mid, and late century for all four seasons). Table 16.—Projected changes in annual mean, minimum, and maximum temperatures and precipitaion in the assessment area averaged over 30-year periods Baseline (1971-2000)a Mean temperature (°F) Annual 51.1 Winter (Dec-Feb) 31.2 Spring (Mar-May) 50.2 Summer (Jun-Aug) 70.1 Fall (Sep-Nov) 53.0 Minimum temperature (°F) Annual 40.0 Winter (Dec-Feb) 21.7 Spring (Mar-May) 38.1 Summer (Jun-Aug) 58.5 Fall (Sep-Nov) 41.6 Maximum temperature (°F) Annual a 62.3 Winter (Dec-Feb) 40.7 Spring (Mar-May) 62.4 Summer (Jun-Aug) 81.6 Fall (Sep-Nov) 64.4 Scenario Departure from baseline 2010-2039 2040-2069 2070-2099 PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI 0.8 2.0 0.7 1.6 0.3 0.8 0.9 3.2 1.4 2.5 1.5 5.3 2.1 4.1 1.3 4.4 1.4 6.9 1.5 5.5 1.9 7.8 2.1 5.5 1.8 7.1 1.8 9.4 1.7 9.0 PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI 0.7 1.9 0.6 1.5 0.5 1.0 0.7 2.8 1.1 2.3 1.4 5.2 2.1 4.4 1.3 4.6 1.4 6.5 0.9 5.2 1.9 7.7 2.3 5.9 1.9 7.1 1.7 9.0 1.5 8.7 PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI PCM B1 GFDL A1FI 0.9 2.1 0.8 1.6 0.0 0.6 1.1 3.6 1.7 2.7 1.7 5.3 2.0 3.9 1.2 4.2 1.3 7.3 2.1 5.8 1.9 7.8 1.9 5.2 1.9 7.2 1.9 9.8 1.9 9.2 The 1971 through 2000 value is based on observed data from weather staions. 91 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES An n u a l 75 70 T e m p e ra tu re (0 F ) 65 60 55 50 45 40 35 1971 - 2000 2010 - 2039 2040 - 2069 2070 - 2099 GFD L A 1FI Mean G F D L A 1 F I M in im u m G F D L A 1 F I M a xim u m P C M B 1 Mean P C M B 1 M in im u m P C M B 1 M a xim u m Figure 24.—Projected changes in annual mean, minimum, and maximum temperatures across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. See Appendix 3 for projected changes by season. Changes in mean temperature are projected to vary greatly by season. Under PCM B1, winter is projected to warm the most by the end of the century (2.1 °F), followed by spring and summer (1.8 °F), and fall (1.7 °F). For GFDL A1FI, greater increases are projected for summer (9.4 °F) and fall (9.0 °F) than spring (7.1 °F) and winter (5.5 °F). Maximum temperatures are projected to increase more than minimum temperatures for both scenarios across nearly all seasons. Winter is the exception to this trend, with minimum temperature projected to increase by 2.3 °F for PCM B1 and by 5.9 °F for GFDL A1FI by the end of the century (Table 1). Maximum annual temperatures are projected to change by 1.9 °F for PCM B1 and 7.8 °F for GFDL A1FI by the end of the century. 92 These changes in temperature are projected to differ across the assessment area (Figs. 25 through 27). For example, the Ohio portion is projected to experience larger mean and minimum temperature increases during winter at the end of the century for both scenarios than other locations in the assessment area. The Ohio portion is also projected to experience larger end-of-century increases in maximum temperature during summer. This pattern holds true for early- and mid-century projections in the Ohio portion, with the addition that fall maximum temperature during these periods is also projected to increase (Appendix 3). There are also noticeable areas within the higher-elevation Allegheny Mountains in West Virginia and Maryland (Section M221B) that are projected to cool slightly ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Figure 25.—Projected diference in daily mean temperature at the end of the century (2070 through 2099) compared to baseline (1971 through 2000) for two climate scenarios. 93 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Figure 26.—Projected diference in daily minimum temperature at the end of the century (2070 through 2099) compared to baseline (1971 through 2000) for two climate scenarios. 94 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Figure 27.—Projected diference in daily maximum temperature at the end of the century (2070 through 2099) compared to baseline (1971 through 2000) for two climate scenarios. 95 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES under PCM B1, and warm by several degrees less than lower-elevation areas under GFDL A1FI. The baseline climate (1971 through 2000) of this section is consistently several degrees cooler than surrounding areas (see Chapter 3: Fig. 15), and temperatures at the end of the century are projected to be several degrees cooler. A narrow strip running parallel to the southwest to northeast ridges in this section is projected to be 1 to 3 °F cooler for PCM B1 at the end of the century compared to the baseline climate. This trend is visible for mean, minimum, and maximum temperatures through all seasons. The pattern is also visible for GFDL A1FI, which projects warming in this strip, but several degrees less than surrounding areas. Although the two climate scenarios project different amounts of warming, they are largely in agreement that mean, maximum, and minimum temperatures will increase throughout much of the assessment area both annually and in all seasons. The two models are less in agreement about projections of seasonal change, with PCM B1 projecting winter temperature to increase the most (1.8 °F increase in mean temperature) and GFDL A1FI projecting summer and fall to increase the most (8.5 °F and 8.1 °F, respectively). See also Box 11. Box 11. Revisiing the “Warming Hole” In Chapter 3, we discussed the “warming hole” that has been observed across the central United States. Although the core of the warming hole is centered on Midwestern states, the efect extends into the assessment area to a lesser degree, characterized by a reducion in summer high temperatures over the past several decades. Will this patern coninue into the future? If we examine only the staisically downscaled GCM data presented in this chapter, we might conclude that the warming hole will be gone in the next century. However, at least one study suggests that the large grid-scale of GCMs fails to account for regional-scale processes that are important contributors to the warming hole (Liang et al. 2006). Using a dynamical downscaling approach to compare a ine-scale (18.6 miles) regional climate model, CMM5, with the PCM model as an input, this study found a large discrepancy between the downscaled projecions and the coarse-scale PCM projecions in the central 9 United States. Although both projected an increase in summer temperature, the downscaled CMM5 projected an increase of less than 0.5 °F, whereas the coarse-scale PCM projected a mid-century increase of 5.4 °F or more averaged over 10 years (2041 through 2050). The staisically downscaled projecions for PCM presented in this chapter also suggest a more modest mid-century (2040 through 2069) increase of 0.5 °F in mean summer temperature. So what does this mean for the “warming hole”? These results suggest that, as with past observaions, there may coninue to be regional climate processes, such as cumulus cloud formaion, that reduce the amount of warming experienced during the summer months, at least over the short term. However, dynamical downscaling studies such as this one remain limited, further jusifying the consideraion of a range of potenial future climate scenarios when preparing for future climate change. ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Precipitaion Due to the highly variable nature of precipitation and difficulty in modeling it, projections of precipitation differ considerably from model to model, and generally carry with them a higher level of uncertainty than projections of temperature (Kunkel et al. 2013b, 2013c; Winkler et al. 2012). The two climate model-scenario combinations used in this assessment describe a wide range of possible future precipitation for the assessment area (Figs. 28 and 29). However, other GCM and emissions scenario combinations could project values outside of this range. Within the assessment area, annual precipitation is projected to increase by 2 inches for PCM B1 and only slightly (0.2 inch) for GFDL A1FI at the end of the century (Table 17) (see Appendix 3 for maps of projected changes in early- and mid-century precipitation). It is more important, however, to consider changes by season, as the timing of increases or decreases have the most implications for forest ecosystems. Under PCM B1, precipitation is projected to increase in winter (0.7 inch), spring (0.7 inch), and summer (1.8 inches) and decrease in fall (-1.2 inches). Under GFDL A1FI, precipitation is projected to increase in fall (0.4 inch), winter (2.1 inches), and spring (1.7 inches) and decrease in summer (-4.1 inches). Notably, for GFDL A1FI, an increase of 1.7 inches in spring precipitation is followed by a decrease of 4.1 inches in summer precipitation at the end of the century. That represents a 13-percent increase from baseline precipitation (Chapter 3) in spring, followed by a 48-percent decrease in summer. These projected summer and fall decreases in precipitation, and their timing during the growing season, could have important consequences for tree growth, seedling establishment, and other forest processes that are dependent on adequate soil moisture. An n u a l 46 P re c ip ita tio n (in c h e s ) 45 44 43 42 1971 - 2000 2010 - 2039 GFD L A 1FI 2040 - 2069 2070 - 2099 PCM B1 Figure 28.—Projected changes in annual mean precipitaion across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. See Appendix 3 for projected changes by season. 97 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Figure 29.—Projected diference in mean precipitaion at the end of the century (2070 through 2099) compared to baseline (1971 through 2000) for two climate scenarios. 98 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Table 17.—Projected changes in mean precipitaion in the assessment area averaged over 30-year periods Baseline (1971-2000)a Departure from baseline 2040-2069 2070-2099 Scenario 2010-2039 43.1 PCM B1 GFDL A1FI 0.2 -0.3 1.1 -1.1 2.0 0.2 Winter (Dec-Feb) 9.2 PCM B1 GFDL A1FI 0.0 0.9 0.6 1.2 0.7 2.1 Spring (Mar-May) 11.5 PCM B1 GFDL A1FI 0.8 0.7 1.0 0.5 0.7 1.7 Summer (Jun-Aug) 12.7 PCM B1 GFDL A1FI 0.7 -1.1 1.4 -2.6 1.8 -4.1 9.7 PCM B1 GFDL A1FI -1.3 -0.7 -2.0 -0.2 -1.2 0.4 Precipitaion (inches) Annual Fall (Sep-Nov) a The 1971 through 2000 value is based on observed data from weather staions. Annual precipitation across the assessment area is projected to increase throughout the 21st century for PCM B1, with the rate of change increasing after the early century time period. Under GFDL A1FI, precipitation is projected to decrease through mid-century before ultimately increasing slightly at the end of the century (Fig. 28). The seasonal precipitation trends for summer and fall exhibit even more departure from the baseline between the two scenarios (Appendix 3). For example, PCM B1 projects summer precipitation to increase steadily through the end of the 21st century, but GFDL A1FI projects summer precipitation to steadily decrease. Projections for fall follow a similar pattern, but the magnitude of change is less. These changes in precipitation are projected to vary across the assessment area (Fig. 29). Similar to differences in past and future temperature, there is a noticeable trend of decreased precipitation that corresponds with the higher-elevation Allegheny Mountains in West Virginia and Maryland (Section M221B). The baseline climate (1971 through 2000) of this section is consistently much wetter than surrounding areas (see Chapter 3: Fig. 15), especially in spring and summer. Precipitation at the end of the century, however, is projected to decrease more than surrounding areas, by as much as 4 to 5 inches for both PCM B1 and GFDL A1FI. This trend is visible through all seasons. Precipitation is also projected to vary spatially by season, notably with a projected summer increase followed by the fall decrease for PCM B1. Under GFDL A1FI, this sign change occurs earlier in the season, with a projected spring increase followed by summer decrease. 99 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES PRoJECTED TRENDS iN EXTREMES Mean temperature and precipitation are not the only climatic factors that can have important effects on forest ecosystems. As outliers from the average climate, extreme weather events are difficult to forecast and model reliably. In general, there is less confidence in model projections of extreme events over the next century compared with general temperature and precipitation changes, but recent research is beginning to provide more evidence for the magnitude and direction of change in many extreme weather events across the eastern United States (Kunkel et al. 2013a). Extreme Temperatures In addition to projecting mean temperatures, downscaled daily climate data can be used to estimate the frequency of extreme high and low temperatures in the future. Studies of extreme temperatures often define hot days as days hotter than 95 °F and cold days as days colder than 32 °F. A study of the United States projects an increase in hot days in the next three decades (Diffenbaugh and Ashfaq 2010). However, heat waves are difficult to analyze regionally because a heat wave in one area may be considered within the normal temperature range in another area. To account for anomalies across a broad landscape, temperature extremes are often analyzed using the distribution of temperatures (e.g., 95th percentile of maximum daily temperature) or a specific threshold temperature (e.g., 95 °F). Studies from across the Midwest and Northeast consistently project 20 to 30 more hot days per year by the end of the century (Diffenbaugh et al. 2005, Ebi and Meehl 2007, Gutowski et al. 2008, Intergovernmental Panel on Climate Change [IPCC] 2012, Meehl and Tebaldi 2004, Winkler et al. 2012). Under the A2 emissions scenario (see Chapter 2), the West Virginia and Maryland portions of the assessment area are projected to double their number of hot days by 2050 (Horton et al. 2013). The number of days above 90 °F is projected to increase 100 by 19 days in the Midwest and 2 days in the Northeast by mid-century, and days over 100 °F are projected to increase by 11 and 8 days, respectively (Kunkel et al. 2013b, 2013c). Furthermore, the hottest days that occur every 20 years are projected to occur every other year by the end of the century (Gutowski et al. 2008). The frequency of multi-day heat waves is also projected to increase by 3 to  days in southeastern Ohio and northwestern West Virginia (Diffenbaugh et al. 2005). The frequency of cold days and cold nights in the assessment area is projected to decrease by 12 to 15 days by the end of the century (Diffenbaugh et al. 2005). These trends are consistent with assessments covering the entire Midwest and Northeast regions, which projected that the assessment area could experience 22 to 2 fewer days below 32 °F and 9 to 10 fewer days below 0 °F by the middle of the 21st century (Kunkel et al. 2013b, 2013c). Intense Precipitaion As described in Chapter 3, there is a clear trend toward more heavy precipitation events in the Midwest and Northeast (Gutowski et al. 2008, Kunkel et al. 2008, Saunders et al. 2012). Rainfall from these high-intensity events represents a larger proportion of the total annual and seasonal rainfall, meaning that the precipitation regime is becoming more episodic. Climate models project an overall increase in the number of heavy precipitation events globally by the end of the century (IPCC 2007, 2012). Global model projections indicate a potential increase in these events in the central and northeastern United States, especially during winter months (IPCC 2012). Future climate projections for the contiguous United States indicate that the Central Appalachians may experience 2 to 4 more days of heavy (greater than 3 inches) precipitation annually by the end of the century (2070 through 2095) (Diffenbaugh et al. 2005). The same study projected that the frequency of dry days will increase by 8 to 10 days annually by the end of the century. ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Multiple models originating from the Climate Model Intercomparison Project (15 models), statistically downscaled models (8 models), and dynamically downscaled models (11 models) were run under a high emissions scenario (A2) and a low scenario to create a range of simulations for comparison of projections of precipitation and extremes. Multiple simulations for the Midwest (including the Ohio portion of the assessment area) generally agree that mid-century heavy precipitation days (greater than 1 inch) could increase by 10 to 20 percent, although models differ widely (Kunkel et al. 2013b). Downscaled projections for the Northeast (including the West Virginia and Maryland portions of the assessment area) indicate increases of up to 30 percent in heavy precipitation events (Kunkel et al. 2013c). Within some areas in West Virginia, more than 50 percent of climate models show increases. Although simulations consistently project an upward trend in extreme events, the magnitude of change is more uncertain, reflecting the high spatial and temporal variability in extreme precipitation data. It is important to consider this trend in combination with the projected changes in mean precipitation over the 21st century. A given increase or decrease in precipitation is unlikely to be distributed evenly across a season or even a month. Additionally, large-scale modeling efforts have also suggested that climate change will increase the year-to-year variability of precipitation across the Midwest and Northeast (Boer 2009). Further, ecological systems are not all equally capable of holding moisture that comes in the form of extreme events. Areas dominated by very coarse- or very fine-textured or shallow soils may not have the water holding capacity to retain moisture received during intense rainstorms. More episodic rainfall could result in increased risk of drought stress between rainfall events or higher rates of runoff during rainfall events. Landscape position will also influence the ability of a particular location to retain moisture from extreme events; steep slopes shed runoff faster than flatter surfaces. Severe Weather: Thunderstorms, hurricanes, and Tornadoes The frequency of strong convective storms has increased in recent decades over the entire Midwest region (Changnon 2011a, 2011b; Diffenbaugh et al. 2008). Projected changes in temperature, precipitation, and convective available potential energy are expected to result in more frequent days over the next century with conditions that are favorable for severe storms (Trapp et al. 2007, 2009, 2011). Several model simulations project increases in thunderstorm frequency within the assessment area for both mid-range (A1B) and higher (A2) emissions scenarios (Trapp et al. 2007, 2009). These changes in storm-forming factors are also expected to influence the formation of tornadoes, although a recent synthesis report on extreme weather events stated that “there is low confidence in projections of small spatial-scale phenomena such as tornadoes and hail because competing physical processes may affect future trends and because current climate models do not simulate such phenomena” (IPCC 2012). As the sophistication of global and regional climate models increases, our understanding of how patterns in hail and tornadoes may change in the future will as well. A recent study using five model simulations projected that the frequency of days favorable for tornadoes rated F2 and greater will increase, and that the peak of tornado season may shift earlier in the season, from May to April (Lee 2012). Projections of hurricane frequency have been associated with too much uncertainty for identifying a clear trend, but it is likely that the spatial distribution of hurricanes will change (Gutowski et al. 2008). For every 1.8 °F increase in sea surface temperature, North Atlantic hurricanes are expected to develop increased wind speeds (1 to 8 percent) and core rainfall rates ( to 18 percent) (Gutowski et al. 2008). Orographic effects of tropical storms and hurricanes in the mountainous sections of the assessment area also have the potential to increase precipitation and subsequent flooding of river channels (Sturdevant-Rees et al. 2001). 101 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES The Allegheny Mountains, home to a diverse array of high-elevaion wetlands. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. PhYSiCAL PRoCESSES Information regarding how temperature and precipitation patterns may change across the assessment area can further be used to examine how these changes may affect the cycling of water in forest ecosystems. Across the globe, increases in temperature are projected to intensify the hydrologic cycle, leading to greater evaporative losses and more heavy precipitation events (IPCC 2007). By examining soil moisture, evapotranspiration, and various drought indices, we can gain an understanding of how these changes may affect water availability for trees, understory plants, 102 wetlands, and rivers. In addition, examining changes in runoff and streamflow can help us assess potential flood risks and changes in watershed dynamics. Flooding and Streamlow Floods occur from a combination of hydrologic, climatological, and biogeographical conditions. High-intensity rainfall events are linked to both localized flash flooding and widespread regional floods, and their effects depend on soil saturation and stream levels at the time of the event. Earlier in this chapter, we discussed projected increases in annual precipitation, and more importantly, a shift ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES towards more episodic and extreme precipitation events. The amount of precipitation that exceeds soil water-holding capacity is available as runoff, which ultimately determines streamflow. Therefore, streamflow can be used as an indicator of the potential for increased flooding, in the absence of more direct indicators. A study in the Mid-Atlantic region projected that increases in temperature at the end of the century would lead to increased evapotranspiration and an increase in summer and fall water deficit (Moore et al. 1997). Consequently, mean annual streamflow was projected to decrease across the assessment area by 21 percent, with the most dramatic decreases occurring in the fall and winter (Moore et al. 1997). Another study in the Mid-Atlantic region projected that increases in precipitation in winter and spring will result in increased streamflow early in the year, and that decreases in precipitation in summer will result in decreased streamflow late in the year (Neff et al. 2000). Snow and Winter Storms Recent studies across much of the Midwest and Northeast have shown that the ratio of snow to rain is strongly correlated with daily mean air temperature in winter (Feng and Hu 2007, Kunkel et al. 2002). Within the assessment area, it is projected that winter mean temperatures will increase by 2.1 °F for PCM B1 and by 5.5 °F for GFDL A1FI by the end of the century, so that winter precipitation in the form of rain is likely to increase. Global models have projected decreases in snow cover across the mid-latitudes with exceptions at high elevation, such as the Sierra Nevada mountain range in the western United States (Hosaka et al. 2005, Kapnick and Delworth 2013). The highest elevations within the assessment area do not produce similar exceptions in these broad-scale models, which project shorter snow duration and decreased snow-water equivalent (IPCC 2007, Lemke et al. 2007). According to two GFDL models, snowfall in the assessment area is projected to decrease by 20 to 50 percent over the next 70 years (Fig. 30) (Kapnick and Delworth 2013). Regional snow cover is projected to decrease by 1.2 to 4 inches by the end of the century for a mid-range emissions scenario (A1B; see Chapter 2) (Hosaka et al. 2005). These are consistent with projections of decreased snow events, snowpack, and snow duration in the Northeast and Midwest (Campbell et al. 2010, Hayhoe et al. 2007, Kunkel et al. 2013a). In general, warming temperatures may lead to a decrease in the overall frequency of ice storms and snowstorms due to a reduction in the number of days that are cold enough for those events to occur. However, there is research to suggest that snowfall in lake-effect areas may increase over the short term if the necessary conditions are present: reduced ice cover on the Great Lakes must result in increased evaporation from the open water, and winter temperatures must remain cold enough for the movement of increased moisture over the land surface to generate snow (Burnett et al. 2003, Wright et al. 2013). Ice cover has declined in recent years on both Lakes Erie and Michigan (Burns et al. 2005, Wang et al. 2012). Projected increases in air temperatures are expected to drive decreases in ice cover duration and extent on the Great Lakes, potentially allowing more winter evaporation and lake-effect snow (Kling et al. 2003, Wright et al. 2013). Shifts in winter precipitation and temperature are expected to advance the timing of snowmelt runoff earlier into the year (Hodgkins and Dudley 200). The ability of soils to absorb this moisture will depend on land cover, infiltration rates, and the soil frost regime (Eisenbies et al. 2007). If soils are able to absorb and retain more of this moisture, soil moisture could be higher at the outset of the growing season. If this moisture is instead lost to runoff, 103 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Figure 30.—Projected change in annual mean snowfall across the United States over the next 70 years (Kapnick and Delworth 2013). forests in the assessment area could be more likely to enter the growing season without sufficient moisture to sustain them throughout the growing season. Snow Cover and Soil Frost The dynamics of snow and frozen soil can have important implications for water availability at the beginning of the growing season. Winter temperatures are projected to increase across the assessment area for both PCM B1 and GFDL A1FI, especially minimum winter temperatures (see Figures 25 through 27). Snow cover typically insulates forest soils, so reduced snowpack could leave the soil surface more exposed to fluctuations in air temperature (Campbell et al. 2010). The degree of warming, and its effects on snowpack, therefore, is likely to determine the impacts on soil temperature, water infiltration, and spring photosynthesis. There are currently no published studies available that have examined this 104 relationship in the assessment area, but studies from adjacent areas can help us understand potential changes. A study that attempted to integrate these complex trends at the Hubbard Brook Experimental Forest used three climate models (Hadley, GFDL, and PCM) for two scenarios (A1FI and B1) through the year 2100 (Campbell et al. 2010). Four of the six scenarios projected increases in total annual and winter precipitation. Although there are no projected changes in soil frost depth, and only a slight increase in freeze-thaw events, the total number of days of soil frost is projected to decline as a direct result of declining snowpack (Campbell et al. 2010). Therefore, it is likely that warmer winter air temperatures will more than counteract the loss of snow insulation and soil frost will generally be reduced across the assessment area. These projections are generally consistent with studies of snowpack and soil frost in the Midwest (Sinha and Cherkauer 2010). ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Drought and Soil Moisture Changes in soil moisture are largely driven by the balance of precipitation and evapotranspiration, and there is some uncertainty about future precipitation changes, especially in areas of complex topography. Further, projections differ widely among models, and an increase in precipitation (and also soil moisture) is expected during the winter and spring. Conversely, decreases are expected in summer or fall, and late-season droughts may become more frequent and more severe, especially when higher air temperatures increase potential evapotranspiration (Gutowski et al. 2008). Many model simulations have projected an increase in summer drying in the mid-latitudes, indicating increased risk of drought (Gutowski et al. 2008). In a study of the northeastern United States, the frequency of short- (1 to 3 months), medium- (3 to  months), and long-term ( months or longer) drought was projected to increase by 3, 0.4, and 0.04 droughts, respectively, per 30-year time interval (Hayhoe et al. 2007). The Variable Infiltration Capacity model, used to explore seasonal soil saturation across the United States during 2071 through 2100, also projected summer and fall decreases in soil moisture, with the greatest decrease (10 percent) in the West Virginia portion of the assessment area (Ashfaq et al. 2010). These broad-scale trends can be useful for estimating a range of potential changes; however, local soil moisture responses to changes in temperature and precipitation are likely to be highly variable within the Central Appalachians, depending on landscape position, normal variability in weather events, and degree of climate change. Evapotranspiraion Evapotranspiration is an important indicator of moisture availability in an ecosystem and the amount of water available to be lost as runoff. Increased precipitation can provide more water available to be evaporated from the soil or transpired by plants. Increased temperature can also drive increases in evapotranspiration, but only as long as there is enough water available. Projected changes in evapotranspiration differ considerably by hydrologic model and climate models used, and whether changes in vegetation are also considered. A study using a regional climate model to examine changes across the continental United States projected an increase in evapotranspiration across the assessment area in summer, which was closely associated with increased precipitation and soil moisture (Diffenbaugh et al. 2005). Another study examining changes averaged over 2071 through 2100 projected increases in evapotranspiration across the assessment area in spring (Ashfaq et al. 2010). In the summer, the largest increases in evapotranspiration were projected in the Allegheny Mountains. Moderate increases during fall were projected mostly east of the Allegheny Mountains, and there was little to no change in evapotranspiration during winter (Ashfaq et al. 2010). Projections of evapotranspiration were modeled at a finer scale by Pitchford et al. (2012) within the mountainous Mid-Atlantic Highlands region of the assessment area (covering all but the Ohio portion). This study area is topographically complex, with microclimates that are cooler and warmer than regional averages. As temperatures increased by 1.8 and 9 °F, evapotranspiration increased by 0.2 and 1.3 inches per month, with much of the change occurring in the summer months. These results suggest that increasing temperatures could reduce soil water availability. As we will discuss in Chapters 5 and , climate change is further projected to affect the distribution of trees and other plant species, which could also affect evapotranspiration on the landscape. Increases in carbon dioxide are expected to lead to changes in the water use efficiency of vegetation (Drake et al. 1997), but these changes are not currently accounted for in model projections of evapotranspiration across the region. 105 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES Growing Season Length The assessment area has experienced shifts in the growing season over the past century, as noted in Chapter 3. Growing seasons are dictated by a variety of factors, including day length, air temperatures, soil temperatures, and dates of first and last frost (Linderholm 200). Therefore, there are a variety of metrics to describe how growing seasons may continue to change for a range of climate scenarios. A study covering the entire Midwest region (including the Ohio portion of the assessment area) examined the changes in dates for the last spring frost and first fall frost by using two models (PCM and HadCM3) for four climate scenarios (Wuebbles and Hayhoe 2004). This study projected that the growing season will be extended by 30 days for the B1 emissions scenario and 70 days for the A1FI scenarios as the last spring frost dates are projected to shift earlier into the year at approximately the same rate that first fall frost dates will retreat later into the year. A study covering the Northeast (including the West Virginia and Ohio portions of the assessment area) examined changes in the last spring frost and first fall frost by using multiple models with the A2 scenario (which projects lower greenhouse gas emissions than A1FI at mid-century) and predicted that the freeze-free season will increase by 19 days by 2055 (Fig. 31) (Kunkel et al. 2013c). A similar study of the freeze-free season in the Midwest region (including the Ohio portion of the assessment area) projected an increase of 22 to 25 frost-free days (Fig. 31) (Kunkel et al. 2013b). How this translates into the actual length of the growing season, as determined by leaf-out and senescence, has not yet been examined for the region. Figure 31.—Projected changes in length of frost-free season across the Midwest (Kunkel et al. 2013b) and Northeast (Kunkel et al. 2013c). Projecions from 2041 through 2070 are shown relaive to the 1980 through 2000 baseline. Projecions are the mean of eight simulaions for the A2 scenario. Modiied from Kunkel et al. (2013b, 2013c). 10 ChAPTER 4: PRoJECTED ChANGES iN CLiMATE, EXTREMES, AND PhYSiCAL PRoCESSES ChAPTER SuMMARY Projected trends in annual, seasonal, and monthly temperature (mean, minimum, and maximum) and total precipitation indicate that the climate will continue to change through the end of this century. Temperatures are projected to increase across all seasons, with extreme warming for the high emissions scenario over the 21st century. The “worst-case scenario” (A1FI) projects annual temperatures that reach 8 to 10 °F higher than the last 30 years of the 20th century. The PCM B1 scenario, despite projecting only slight increases for other seasons, projects winter minimum temperature to increase by 2 to 4 °F over most of the assessment area. Precipitation is projected to increase in winter and spring by 2 to 5 inches (depending on scenario), leading to potential spring increases in runoff and streamflow. Projections of precipitation differ among climate models in summer and fall; however, higher temperatures during those seasons mean that much of that precipitation will contribute to increased evapotranspiration. Changes in temperature and precipitation are projected to lead to changes in extreme weather events and local hydrology. There is fairly high certainty that heavy precipitation events will increase, snow cover will decrease, and eventually soil frost will decrease as well. However, more uncertainty remains with respect to changes in tornadoes and thunderstorms, seasonal soil moisture patterns, and flooding. In the next chapter, we examine the ecological implications of these anticipated changes on forest ecosystems. 107 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Changes in climate have the potential to profoundly affect forests of the Central Appalachians region. Many tree species that are currently present may fare worse with warmer temperatures and altered precipitation patterns. Other species may do better under these conditions, and some species not currently present may have the potential to do well if conditions allow them to disperse to newly suitable areas. In addition, climate change can have indirect effects on forests in the region by changing the populations and dynamics of insect pests, pathogens, invasive species, nutrient cycling, and wildfire regimes. In this chapter, we summarize the potential impacts of climate change on forests in the Central Appalachians region over the next century, with an emphasis on changes in tree species distribution and abundance. MoDELED PRoJECTioNS oF FoREST ChANGE Forest ecosystems in the assessment area may respond to climate change in a variety of ways. Potential changes include shifts in the spatial distribution, abundance, and productivity of tree species. For this assessment, we rely on a combination of three forest impact models to describe these potential changes: the Climate Change Tree Atlas (DISTRIB), LINKAGES, and LANDIS PRO (Table 18). The Tree Atlas uses statistical techniques to model changes in suitable habitat for individual species over broad geographic areas. LINKAGES predicts establishment and growth of trees based on soils and other site information and climate. LANDIS PRO simulates changes in basal area and trees per acre to project the abundance and 108 distribution of individual tree species. No single model offers a comprehensive projection of future impacts on forest ecosystems, but each tool is valuable for a particular purpose or set of questions. Similarities in patterns across models suggest less uncertainty in projections than when patterns differ, and differences in patterns provide opportunities to better understand the nuances of ecological responses given the strengths and limitations of the models. For a more thorough description of the different models, and specifically how they were applied for this assessment, see Chapter 2. All three research teams used the same downscaled climate projections from two combinations of general circulation models (GCMs) and emissions scenarios described in detail in Chapter 4: GFDL A1FI and PCM B1. Projected changes in temperature and precipitation for GFDL A1FI represent the higher end of the range of changes, and projections for PCM B1 represent the lower end. This consistency in the climate data used in each modeling approach means that the forest impact models describe potential forest changes over the same range of plausible future climates. These model results are most useful to describe trends across large areas and over long time scales. These models are not designed to deliver precise results for individual forest stands or a particular year in the future, despite the temptation to examine particular data points or locations on a map. Maps are spatially representative but not spatially exact. In this chapter, we present model results for the end of the 21st century. Data for intermediate time periods are provided in Appendix 4. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Table 18.—Overview of the three models used in this assessment Feature Tree Atlas LiNKAGES LANDiS PRo Summary Suitable habitat distribuion model (DISTRIB) + supplementary informaion (modifying factors) Patch-level forest succession and ecosystem dynamics process model Spaially dynamic forest landscape process model Primary outputs for this assessment Area-weighted importance values and modifying factors by species Species establishment and growth maps (% change) Basal area, and trees per acre by species Model-scenario combinaions GFDL A1FI and PCM B1 Assessment area Central Appalachians assessment area within Ohio, West Virginia, Maryland: Secions 221E and 221F in Province 221; Secions M221A, M221B, and M221C in Province M221 Resoluion 12-mile grid 0.2-acre (1/12-ha) plots represening landforms in subsecions 886-foot grid Number of species evaluated 94 23 17 Control/baseline climate 1971 through 2000 1990 through 2009 n/a Climate periods evaluated 2010 through 2039, 2040 through 2069, 2070 through 2099 1990 through 2009, 2080 through 2099 2009 through 2099 Simulaion period n/a 30 years 2009 through 2099 Compeiion, survival, and reproducion No Yes Yes Disturbances No (but addressed through modifying factors) No Timber harvest Tree physiology feedbacks No Yes No Succession or ecosystem shits No No Yes Biogeochemical feedbacks No Yes No 109 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Tree Atlas Importance values of 134 eastern tree species were modeled for potential habitat suitability in the assessment area by using the DISTRIB model, a component of the Tree Atlas toolset (Landscape Change Research Group 2014). From U.S. Forest Service Forest Inventory and Analysis (FIA) data, the number of stems and the basal area were used to calculate importance values for each 12.4-mile grid cell for each tree species. For an individual grid cell, the importance value for a species can range from 0 (not present) to 100 (completely covering the area). Importance values for each pixel were then summed across the assessment area to calculate the area-weighted importance value for a species; thus area-weighted importance values can be greater than 100. This analysis was conducted for the assessment area and for individual ecological sections within the assessment area. Results for the entire assessment area are presented in the text below. Appendix 4 contains the full set of results summarized by ecological section. More information on Tree Atlas methods can be found in Chapter 2. Of the 134 species modeled, 93 currently have or are projected to have suitable habitat in the area. The projected changes in suitable habitat for the 93 species were calculated for the years 2070 through 2099 for the GFDL A1FI and PCM B1 scenarios and compared to present values (Table 19). Species were categorized based upon whether the results from the two climate scenarios projected an increase, decrease, or no change in suitable habitat compared to current conditions, or if the model results were mixed. Further, some tree species that are currently not present in the assessment area were identified as having potential suitable habitat in the future for one or both scenarios. See Appendix 4 for complete results from the DISTRIB model for early (2010 through 2039), middle (2040 through 209), and end (2070 through 2099) of century time periods. Roughly half of the tree species modeled 110 are found in every section of the assessment area, whereas half are missing from at least one section. Section M221B (Allegheny Mountains) contains the highest number of species (74), and Section M221C (Northern Cumberland Mountains) has the lowest number of species (2). This is not an accurate reflection of species diversity, however, because only the most common species were modeled. Modifying factors have also been incorporated into the Tree Atlas to provide additional information on potential forest change. Modifying factors include life history traits and environmental factors that make a species more or less able to persist in the eastern United States (Matthews et al. 2011b). These factors are not explicitly included in the DISTRIB outputs, and are based on a review of a species’ life-history traits, known stressors, and other factors. Examples of modifying factors include drought tolerance, dispersal ability, shade tolerance, site specificity, and susceptibility to insect pests and diseases. Factors are identified for a species throughout its range and do not account for sitespecific conditions which may also influence a species’ potential for change. For each modifying factor, a species was scored on a scale from -3 (very negative response) through +3 (very positive response), and further weighted by confidence and relevance to future projected climate change. The means of these scores were plotted to determine an overall score of adaptability (see Appendix 4 for detailed methods). Information on adaptability is included in the summary of projected changes in habitat (Table 19), where a plus (+) or minus (-) sign after a species name indicates that certain modifying factors could lead the species to do better or worse, respectively, than DISTRIB model results indicate. As an example, the species with the five highest and five lowest adaptability scores are displayed in Table 20. Appendix 4 contains more information on the specific modifying factors and overall adaptability scores for each species. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Table 19.—Potenial changes in suitable habitat for 93 tree speciesa in the Central Appalachians region Common name PCM B1 GFDL A1Fi Small Decrease Large Decrease Large Decrease Small Decrease Small Decrease Small Decrease Large Decrease Large Decrease Small Decrease Small Decrease Large Decrease Exirpated Small Decrease Large Decrease Exirpated Large Decrease Large Decrease Large Decrease Large Decrease Large Decrease No Change under Both Scenarios American chestnut American holly American hornbeam Bear oak (scrub oak) Blackgum (+) Cucumbertree Mountain maple (+) Northern pin oak (+) Pignut hickory Pitch pine Serviceberry Southern magnolia Tamarack (naive) (-) Yellow buckeye (-) No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change No change increases under Both Scenarios Biternut hickory (+) Blackjack oak (+) Chinkapin oak Common persimmon (+) Eastern redcedar Eastern redbud Green ash Hackberry (+) Honeylocust Osage-orange (+) Post oak (+) Shagbark hickory Shingle oak Shortleaf pine Southern red oak Sugarberry Sweetgum Winged elm Large Increase Small Increase Large Increase Large Increase Large Increase Small Increase Small Increase Small Increase Small Increase Small Increase Large Increase Small Increase Small Increase Large Increase Large Increase Large Increase Large Increase Large Increase PCM B1 GFDL A1Fi Mixed Results Declines under Both Scenarios Balsam ir (-) Bigtooth aspen Black ash (-) Black cherry (-) Chokecherry Pin cherry Quaking aspen Red pine Striped maple Yellow birch Common name Large Increase Large Increase Large Increase Large Increase Large Increase Small Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase Large Increase American basswood No Change American beech No Change American elm No Change Black locust No Change Black maple No Change Black oak No Change Black walnut Small Increase Black willow (-) Small Decrease Blue ash (-) No Change Boxelder (+) No Change Bur oak (+) No Change Buternut (-) No Change Chestnut oak (+) No Change Eastern cotonwood Small Decrease Eastern hemlock (-) No Change Eastern hophornbeam (+) No Change Eastern white pine No Change Flowering dogwood No Change Loblolly pine (-) No Change Mockernut hickory No Change Northern catalpa No Change Northern red oak No Change Ohio buckeye (-) No Change Pawpaw No Change Pin oak (-) No Change Red maple (+) No Change Red mulberry No Change Red spruce (-) No Change River birch No Change Rock elm (-) No Change Sassafras No Change Scarlet oak No Change Shumard oak (+) NA Silver maple (+) Small Decrease Slippery elm No Change Sourwood (+) Small Increase Sugar maple (+) No Change Swamp white oak No Change Sweet birch No Change Sycamore No Change Table Mountain pine (+) No Change Tulip tree No Change Virginia pine Small Increase White ash (-) No Change White oak (+) No Change Willow oak NA Large Decrease Large Decrease Small Decrease Small Decrease Large Decrease Large Increase No Change Large Increase Small Decrease Small Increase Large Increase Exirpated Small Decrease Large Increase Small Decrease Small Increase Large Decrease Small Decrease Large Increase Small Increase Small Increase Small Decrease Large Decrease Large Decrease Small Increase Large Decrease Large Increase Large Decrease Small Increase Large Increase Small Decrease Small Decrease Large Increase Large Increase Small Decrease Small Decrease Large Decrease Large Decrease Large Decrease Small Increase Small Increase Large Decrease No Change Large Decrease Small Increase Large Increase (coninued on next page) 111 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Table 19 (coninued). Common name PCM B1 GFDL A1F New Habitat NA NA NA NA New Habitat New Habitat New Habitat New Habitat New Habitat New Suitable habitat Black hickory Cedar elm Northern white-cedar Water locust Water oak Species are grouped according to change classes (e.g., increase, no change) based on the percentage change in the areaweighted importance value projected for the end of century (2070 through 2099) for two climate-emissions scenarios. Species with the 20 highest or 20 lowest modifying factor scores are marked with plus (+) and minus (-) signs, respecively. Appendix 4 contains descripions of change classes and complete results for all 93 species for the assessment area and for each ecological secion. a When examining these results, it is important to keep in mind that model reliability is generally higher for common species than for rare species. FIA data also tend to undersample riparian areas as they are usually narrow strips within an upland matrix (Iverson et al. 2008). When model reliability is low, less certainty exists for the model results. See Appendix 4 for specific rankings of model reliability for each species. Table 20.—Species with the ive highest and ive lowest values for adapive capacity based on Tree Atlas modifying factors Species Factors that afect raing Highest adapive capacity 1. Red maple high probability of seedling establishment, wide range of habitats, shade tolerant, high dispersal ability 2. Boxelder high probability of seedling establishment, shade tolerant, high dispersal ability, wide range of temperature tolerances, drought tolerant 3. Sourwood good light compeitor, wide range of habitats 4. Bur oak drought tolerant, ire tolerant 5. Eastern hophornbeam shade tolerant, wide range of temperature tolerances, wide range of habitats Lowest adapive capacity 1. Black ash emerald ash borer suscepibility, poor light compeitor, limited dispersal ability, poor seedling establishment, ire intolerant, dependent on speciic hydrologic regime 2. Buternut buternut canker, drought intolerant, ire intolerant, poor light compeitor 3. Balsam ir spruce budworm and other insect pests, ire intolerant, drought intolerant 4. White ash emerald ash borer, drought intolerant, suscepible to ire topkill 5. Blue ash emerald ash borer suscepibility, limited dispersal ability, ire intolerant, poor light compeitor, dependent on speciic hydrologic regime 112 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Decreases in Suitable habitat For the Central Appalachians region, 10 of the 93 modeled species are projected to undergo large or small declines in suitable habitat for the full range of projected climate futures. Projected declines in habitat (as measured by a ratio of potential future importance value to current importance value) are more severe for GFDL A1FI than PCM B1 for most of these species. These reductions in suitable habitat do not imply that all or most mature trees will die or the species will be extirpated; rather, these results indicate that these species will be living outside their ideal climatic envelope. As a result, trees living in less suitable habitats may have greater susceptibility to stressors, and may also be at greater risk of regeneration failure. Generally, the changing climate tends to intensify or add to the stresses that may already exist for the species and increases susceptibility to drought, pests, diseases, or competition from other species including invasives. Black cherry is currently abundant within the assessment area, but is projected to decline for both scenarios, more so for GFDL A1FI. The nine other species in this category are much less common on the landscape, and are projected to lose a large portion of suitable habitat for GFDL A1FI. Many of the species are currently near the southern limit of their range in the assessment area or exist as disjunct populations. Balsam fir and red pine are glacial relicts that are currently limited to higher elevations in West Virginia, and the majority of these species’ ranges are much farther north (Hessl et al. 2011, Potter et al. 2010). Black ash distribution in the assessment area is closely tied to the Greenbrier Limestone. Decreases in suitable habitat may be catastrophic for these highly localized populations. Other species are not geographically limited, and are therefore more widespread throughout the assessment area. Bigtooth aspen and chokecherry are relatively widespread throughout the assessment area, and are projected to lose all suitable habitat for GFDL A1FI. Black ash, pin cherry, quaking aspen, striped maple, and yellow birch are currently rare on the landscape, and their suitable habitat is projected to decrease substantially. Balsam fir and black ash also have highly negative modifying factors, suggesting that there are lifehistory traits or disturbance stressors that may cause these species to lose even more suitable habitat than the model results indicate. For example, the expanding presence of emerald ash borer in the assessment area is expected to greatly reduce the importance of black ash in the area; its impact will be much more than the loss that is projected to occur from changing climatic conditions. Modifying factors for balsam fir include susceptibility to balsam woolly adelgid and drought. No Change in Suitable habitat Fourteen species are projected to undergo less than a 20-percent change in suitable habitat for both scenarios. American hornbeam, blackgum, and pignut hickory are currently abundant across the region and their habitat is not projected to decrease or increase substantially. Blackgum has one of the highest adaptive capacity scores, partially because of its fire tolerance and shade tolerance, and it is likely to do better than projected. Serviceberry, pitch pine, cucumbertree, and yellow buckeye are less common on the landscape, and American chestnut, tamarack, mountain maple, scrub oak, and others are extremely rare (Appendix 4). Tamarack and yellow buckeye have several negative modifying factors, including habitat specificity and susceptibility to fire, insect pests, and drought, suggesting these species may fare worse than projected. Mixed Results in Suitable habitat The model results projected different outcomes for PCM B1 and GFDL A1FI for almost half of the species (44 of 93). For 23 of these species, DISTRIB projected that suitable habitat will not change or increase for PCM B1 but will decrease for GFDL A1FI, and one species (butternut) was 113 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS projected to lose all suitable habitat. Many of these species are currently common in the assessment area, including American beech, American elm, black locust, black oak, chestnut oak, flowering dogwood, northern red oak, red maple, sassafras, sugar maple, white ash, white oak, and tulip tree. Chestnut oak, red maple, sugar maple, and white oak all have positive modifying factors that indicate that the species may fare better than the models suggest. This is particularly notable for red maple, which has the greatest strength of positive modifying factors among the 134 species that were assessed across the eastern United States. White ash has negative modifying factors due to emerald ash borer and drought mortality and may have even greater decreases than the model predicts. For 17 species, DISTRIB projected that suitable habitat will increase for GFDL A1FI while not changing substantially for PCM B1. Black oak, eastern hophornbeam, mockernut hickory, and white oak are currently common in the assessment area. The remaining species are relatively infrequent or rare, including boxelder, bur oak, pin oak, and sycamore. These species are more frequently found southwest of the assessment area, and DISTRIB results suggest that suitable habitat will move northeast for future conditions. Boxelder, bur oak, and white oak have positive modifying factors, suggesting these species may do better than projected. The remaining six species are projected to change for PCM B1 but have the opposite direction of change or no change for GFDL A1FI. Eastern cottonwood, black willow, and silver maple are projected to lose suitable habitat for PCM B1, but gain suitable habitat for GFDL A1FI. For sourwood, the trend is reversed. Virginia pine and black walnut gain suitable habitat for PCM B1, but maintain their current relative amounts for GFDL A1FI. These species may take advantage of increased temperatures by colonizing habitat at higher 114 elevations that were previously too cool. Sourwood is currently a common species in the assessment area, frequently associated with pine and oak forests. The positive modifying factors associated with this species, including shade tolerance and an ability to occupy a wide range of sites, suggest that it may fare better than what the model projects. increases in Suitable habitat Suitable habitat for 18 species is projected to increase for both models by the end of the century. All of these species are considered rare in the assessment area (see Appendix 4 for Tree Atlas rules regarding rare species). Many of these species, such as blackjack oak, chinkapin oak, common persimmon, eastern redcedar, eastern redbud, hackberry, honeylocust, post oak, shingle oak, shortleaf pine, southern red oak, sweetgum, and winged elm are close to the northern or eastern extent of their range. Several species common to the south of the assessment area may become more widespread throughout the landscape, assuming higher regeneration success for future forest conditions. Because many of the species projected to lose suitable habitat are still expected to be major components of forest ecosystems by the end of the century, forests in the assessment area may have the potential to contain a higher diversity of species in the future, with a blend of southern and temperate species. A few species within the increase category, such as shortleaf pine and winged elm, have negative modifying factors, which suggest that they may be less able to take advantage of increases in suitable habitat. At the same time, several species have positive modifying factors, such as bitternut hickory, blackjack oak, common persimmon, and post oak, and may be better able to cope with potential changes in climate, beyond what the models suggest. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS New Suitable habitat The DISTRIB model projects gains in suitable habitat for five species (black hickory, cedar elm, northern white-cedar, water locust, and water oak) that are currently not present in the assessment area for GFDL A1FI. This result does not necessarily mean that a given species will be able to migrate to newly available habitat and colonize successfully; it indicates instead that conditions may be suitable for a species to occupy the site if it can establish. Habitat fragmentation and the limited dispersal ability of seeds could hinder the northward movement of southern species, despite the increase in habitat suitability (Ibáñez et al. 2008), and most species can be expected to migrate more slowly than their habitats will shift (Iverson et al. 2004a, 2004b). Geographic Trends Projected changes are not uniform across the assessment area, and areas of suitable tree habitat are governed by soils, moisture gradients, and other factors in addition to climate. The geographic and biological complexity of the Central Appalachians region warranted a closer look at the five ecological sections within the broader assessment area (see Chapter 1 for a map). Furthermore, Section 221E was split along the Ohio and West Virginia state lines and Tree Atlas results are available for those two smaller areas. Slightly more than half of the species modeled currently have or are projected to have suitable habitat in all six sections, and onequarter of the species are modeled in four or five sections. The remaining one-quarter of the species are modeled in three or fewer sections. Among the species projected to have suitable habitat across four or more sections of the assessment area, distinct differences in climate, landform, and other characteristics often result in a variety of projected change classes between sections for a single species. Species showing significant geographic trends include: American elm, American hornbeam, blackgum, blackjack oak, black walnut, black willow, boxelder, butternut, chestnut oak, cucumbertree, eastern hophornbeam, eastern white pine, flowering dogwood, green ash, Ohio buckeye, pignut hickory, pitch pine, sassafras, scarlet oak, silver maple, and slippery elm. Appendix 4 shows a comparison of model results by section. Outputs from DISTRIB can also be visualized as maps, such as those available online through the Climate Change Tree Atlas Web site (www.nrs. fs.fed.us/atlas), and these maps can provide greater context for interpreting the projected changes in suitable habitat. It is important to note that these maps detect relative change on a more detailed pixel by pixel basis rather than averaged by section within the assessment area, as presented above. For this assessment, the section boundaries were added to regional Tree Atlas maps in order to help orient the reader. Maps for four species (chestnut oak, sugar maple, eastern white pine, and red spruce) are shown below. Chestnut oak is projected to retain a large amount of suitable habitat in the assessment area for PCM B1, whereas suitable habitat decreases more for GFDL A1FI, with the greatest loss of suitable habitat projected in Section 221F (Fig. 32). Under PCM B1, chestnut oak is projected to gain new suitable habitat in the western portion of the assessment area (221F and 221E Ohio), with no change in the eastern portion (Fig. 32). Sugar maple suitable habitat is projected to decrease across the assessment area, especially around the center of the assessment area (221E Ohio and West Virginia, M221C) for PCM B1, with a much greater loss of habitat projected for GFDL A1FI (Fig. 33). Eastern white pine is currently largely absent or of low importance value across most of the assessment area, with declines projected over most of the current habitat for PCM B1, and complete loss of habitat projected across Province 221 (221F and 221E Ohio and West Virginia). The only suitable habitat remaining for white pine for GFDL A1FI is in Province M221, where suitable habitat is projected to stay the same 115 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Figure 32.—Modeled importance values for chestnut oak across the assessment area and the larger geographic region for current climate condiions (top) and projected for the end of the century (2070 through 2099) for the PCM B1 (middle) and GFDL A1FI (botom) climate scenarios, from the Tree Atlas model. Importance values can range from 0 to 100. An importance value of zero (light yellow) indicates that the species is not present currently, or will not have suitable habitat at the end of the century. 11 Figure 33.—Modeled importance values for sugar maple across the assessment area and the larger geographic region for current climate condiions (top) and projected for the end of the century (2070 through 2099) for the PCM B1 (middle) and GFDL A1FI (botom) climate scenarios, from the Tree Atlas model. Importance values can range from 0 to 100. An importance value of zero (light yellow) indicates that the species is not present currently, or will not have suitable habitat at the end of the century. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS in M221B and M221C, and decrease in M221A (Fig. 34). Red spruce is even more geographically constrained for current climate, largely limited to Section M221B (eastern West Virginia), and habitat suitability is projected to decrease for GFDL A1FI (Fig. 35). Red spruce is currently recovering from past harvesting and fire disturbance, and is expanding on the landscape to refill its niche (Nowacki et al. 2009, Seidel et al. 2009). Earlycentury increases in red spruce due to succession and planting efforts may help the species do better than models project in the short term. These maps should be interpreted carefully. As mentioned above, DISTRIB results indicate only a change in suitable habitat, not necessarily that a given species will be able to migrate to newly available habitat. Additionally, these results do not incorporate the influence of modifying factors (positive for sugar maple and chestnut oak, negative for eastern white pine and red spruce). Suitable habitat maps assessing the whole eastern United States are available online through the Climate Change Tree Atlas Web site (www.nrs.fs.fed.us/atlas) for all the species in this assessment (see Appendix 4). As is the case for interpreting any spatial model outputs, local knowledge of soils, landforms, and other factors is necessary to determine if particular sites may indeed be suitable habitat for a given species in the future. These maps serve only as an illustration of broad trends. The Allegheny Mountains, home to a diverse array of high-elevaion wetlands. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 117 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Figure 34.—Modeled importance values for eastern white pine across the assessment area and the larger geographic region for current climate condiions (top) and projected for the end of the century (2070 through 2099) for the PCM B1 (middle) and GFDL A1FI (botom) climate scenarios, from the Tree Atlas model. Importance values can range from 0 to 100. An importance value of zero (light yellow) indicates that the species is not present currently, or will not have suitable habitat at the end of the century. 118 Figure 35.—Modeled importance values for red spruce across the assessment area and the larger geographic region for current climate condiions (top) and projected for the end of the century (2070 through 2099) for the PCM B1 (middle) and GFDL A1FI (botom) climate scenarios, from the Tree Atlas model. Importance values can range from 0 to 100. An importance value of zero (light yellow) indicates that the species is not present currently, or will not have suitable habitat at the end of the century. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS LiNKAGES The LINKAGES model was used to predict tree growth (biomass) for 23 species within the assessment area after 30 years of establishment and growth on a plot from bare ground (Chapter 2) (Wullschleger et al. 2003). Section-level estimates were derived from the weighted average of 0.2acre plots within  landforms in 2 subsections. We report projected tree growth (biomass) for current climate (1990 through 2009) and projected climate (2080 through 2099) using the PCM B1 and GFDL A1FI climate scenarios. Changes in biomass for PCM B1 and GFDL A1FI are calculated as the difference from projected biomass for a current climate scenario at the end of the century (2080 through 2099). The potential change in biomass for GFDL A1FI and PCM B1 is presented as classes of change in maps for each species (Figs. 3 and 37; Appendix 4). Figure 36.—Projected changes in establishment and growth for sugar maple at the end of the century (2080 through 2099) for two climate scenarios. Figure 37.—Projected changes in establishment and growth for loblolly pine at the end of the century (2080 through 2099) for two climate scenarios. 119 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Species establishment probabilities (SEPs) are an important input into the LANDIS PRO model, and were calculated by standardizing the LINKAGES biomass projections across species on a scale from 0 to 1. Absolute and percentage changes in SEP were calculated for the entire assessment area and are presented in Table 21. Results for each of the five sections within the assessment area are provided in Appendix 4. Species establishment probabilities reflect the ability of a species to establish and grow on a site and can be thought of as a measure of habitat suitability, but they do not account for the effects of interspecific competition and disturbance. Because SEPs are derived from the LINKAGES estimates of biomass, the percentage change in SEPs is largely congruent with the mapped changes in biomass; minor differences exist for some species due to the rescaling and rounding involved in calculating the SEPs. Table 21.—Projected changes in mean species establishment probability (SEP) valuesa from current climate at year 2100 across the assessment area Species American beech American elm Balsam irb Black cherry Blackgum Black oak Chestnut oak Eastern redcedar Eastern hemlock Eastern white pine Flowering dogwood Loblolly pine Northern red oak Pignut hickory Post oak Red maple Red spruceb Scarlet oak Shortleaf pine Sugar maple Tulip tree White ash White oak a Current climate SEP SEP 0.22 0.14 0.00 0.28 0.16 0.19 0.20 0.20 0.13 0.34 0.05 0.13 0.29 0.35 0.05 0.31 0.00 0.17 0.08 0.51 0.76 0.43 0.32 0.22 0.16 0.00 0.30 0.19 0.20 0.20 0.24 0.13 0.35 0.08 0.29 0.31 0.37 0.10 0.34 0.00 0.17 0.22 0.50 0.83 0.53 0.35 Future climate PCM B1 GFDL A1Fi % change SEP % change 0.0 14.3 0.0 7.1 18.8 5.3 0.0 20.0 0.0 2.9 60.0 123.1 6.9 5.7 100.0 9.7 0.0 0.0 175.0 -2.0 9.2 23.3 9.4 SEP absolute values were rounded to two decimal places. Absolute values are small enough that percentage changes are more important. b 120 0.02 0.18 0.00 0.30 0.21 0.13 0.12 0.26 0.01 0.04 0.10 0.51 0.18 0.36 0.17 0.37 0.00 0.04 0.35 0.05 0.86 0.41 0.35 -90.9 28.6 0.0 7.1 31.3 -31.6 -40.0 30.0 -92.3 -88.2 100.0 292.3 -37.9 2.9 240.0 19.4 0.0 -76.5 337.5 -90.2 13.2 -4.7 9.4 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Projected changes in both biomass and SEPs represent a species’ potential growth based on site and climate factors (Chapter 2). Both positive and negative changes in potential growth were consistently greater for GFDL A1FI than PCM B1 (Table 21). Suitable habitat was projected to decrease or be extirpated for PCM B1 for only two species: balsam fir and red spruce. American beech, balsam fir, eastern hemlock, eastern white pine, red spruce, and sugar maple suitable habitat was potentially extirpated from all or portions of the assessment region for GFDL A1FI. Scarlet oak potential growth was projected to decrease across the region for GFDL A1FI. Modest to large increases in potential growth are projected for GFDL A1FI, and to a lesser extent for PCM B1, for American elm, blackgum, eastern redcedar, flowering dogwood, loblolly pine, post oak, shortleaf pine, and tulip tree. Results for other species varied between sections, increasing or decreasing for black cherry, black oak, chestnut oak, northern red oak, pignut hickory, and white ash (Appendix 5). Loblolly and shortleaf pine are projected to have the largest increases in potential growth, partly because their biomass and SEP were very low for current climate and even a small increase could double the biomass. As mentioned above, LINKAGES results indicate only potential growth, not necessarily that a given species will be able to colonize newly available habitat. Some species showing large increases in potential growth are currently absent from the region or have very limited distributions. It would take a long time for them to respond (especially without planting) to this increase in potential growth and establishment. For example, loblolly pine is currently rare and exists mostly in plantations within the assessment area. Future potential growth is projected to increase for both climate scenarios and more so for GFDL A1FI, suggesting that habitat will become more favorable for this southern species. Projected changes in both biomass and SEPs do not represent actual current or future distributions. Furthermore, LINKAGES is not spatially dynamic and does not simulate tree dispersal or any other spatial interaction, such as competition, among grid cells. Rather, this spatial interaction is examined by using LINKAGES results as input in the LANDIS PRO model. LANDiS PRo The LANDIS PRO model was used to simulate changes in basal area (BA) and trees per acre (TPA) for 17 species over 90 years (2009 through 2100). Basal area and number of trees per acre were simulated for each 88-foot grid cell and then summarized for ecological sections and the entire assessment area. The LANDIS PRO model used the SEPs from LINKAGES (see above) to link tree establishment and growth to climate and additional parameters that reflect species life histories and landscape processes such as succession and competition, seed dispersal, and timber harvest. Parameters were initially based on current known silvics for each species, and then adjusted so that simulations using current climate produced values for species abundance that are consistent with FIA data and earlier growth studies in the region. Forest management was simulated as tree harvest on 8 to 13 percent of the forested area per decade, with the older stands harvested first. The model did not include wind or fire disturbance; that is, simulations represent forests with succession and management but without mortality from fire, wind, insects, disease, or other disturbances. The LANDIS PRO model differs substantially from the Climate Change Tree Atlas and LINKAGES because it simulates tree, stand, and landscape dynamics over time; therefore, the composition and structure of a pixel can be examined for any point in time in the 121 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS simulation. Furthermore, LANDIS PRO accounts for natural stand dynamics in addition to climate effects on establishment and growth and is a prediction of actual forest composition and structure for a future year. Future forest composition and structure were reported as BA and TPA for each tree species. Basal area is the area of tree stems at breast height per acre. High BA can be driven by many large-diameter trees, an even greater number of small-diameter trees, or a combination of the two. Therefore, TPA is also included as another measure of abundance and density, regardless of tree size (see Appendix 4 for area graphs of BA and TPA for PCM B1 and GFDL A1FI). A low BA with a high TPA indicates many small trees. A high BA with a low TPA indicates a higher proportion of large trees. Projected change in BA and TPA is presented in area charts for current climate (Figs. 38 and 39) and future climate for PCM B1 and GFDL A1FI through 2100 (Appendix 4). Estimates of percentage change in BA and TPA for current climate were calculated as the change from observed 2009 values and represent change due to succession and management over the time period. Percentage change in BA and TPA for PCM B1 and GFDL A1FI at 2040, 2070, and 2100 was calculated as the change from current climate in the same model year (2040, 2070, or 2100) and represents the change due to the alternative climate, which is in addition to change due to succession and management. Percentage change in BA and TPA at year 2100 is presented in Table 22. Figure 38.—Projected changes in basal area for 17 species across the assessment area under the current climate scenario. Assessment area values were derived from the weighted average of secions. The width of the colored line represents the relaive basal area for each species through the year 2110. For example, red maple had the highest basal area in 2010, and basal area is projected to increase due to natural succession and management for current climate. 122 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Table 22.—Absolute and percentage change in basal area (BA) (square feet per acre) and trees per acre (TPA) predicted by the LANDiS PRo model for 17 species for current and future climate scenarios for the assessment area in year 2100 Tree species BA in 2009 BA by year and climate scenarioa Current PCM B1 BA Change BA Change in from year year from current 2100 2009 2100 climateb GFDL A1Fi BA Change year from current 2100 climateb American beech Black cherry Black oak Chestnut oak Eastern hemlock Eastern redcedar Eastern white pine Loblolly pine Northern red oak Pignut hickory Red maple Red spruce Scarlet oak Sugar maple Tulip tree White ash White oak 5.1 7.3 6.8 5.6 1.7 0.1 1.2 0.1 6.8 2.6 16.0 1.0 4.8 6.1 6.5 2.5 5.7 4.6 4.6 1.5 4.9 1.1 0.0 1.0 0.2 2.8 2.7 30.0 0.3 0.7 8.0 10.5 2.9 8.4 4.2 4.5 1.5 4.6 1.0 0.0 0.9 0.2 3.1 2.7 30.3 0.3 0.7 7.3 10.8 2.9 8.3 Tree species TPA year 2009 TPA by year and climate scenarioa Current PCM B1 TPA Change TPA Change year from year year from current 2100 2009 2100 climateb American beech Black cherry Black oak Chestnut oak Eastern hemlock Eastern redcedar Eastern white pine Loblolly pine Northern red oak Pignut hickory Red maple Red spruce Scarlet oak Sugar maple Tulip tree White ash White oak 26.8 18.9 8.4 14.9 6.1 0.8 7.3 0.5 11.9 7.3 77.4 5.1 4.0 54.1 22.9 11.9 12.8 a 14.2 7.2 2.0 30.9 3.8 0.1 2.5 0.2 3.6 5.4 66.9 0.4 1.1 53.6 104.3 7.3 57.7 -9% -38% -78% -12% -35% -89% -13% 127% -58% 5% 88% -67% -86% 30% 62% 18% 46% -47% -62% -77% 108% -38% -91% -66% -68% -70% -27% -13% -92% -72% -1% 356% -39% 349% 4.6 4.6 1.5 5.1 1.1 0.0 1.0 0.2 3.3 2.8 30.5 0.3 0.7 8.0 10.8 3.0 8.6 13.9 7.3 2.0 32.6 3.6 0.0 2.4 0.2 6.0 5.6 70.3 0.4 1.2 50.8 112.1 7.9 60.9 0% 1% 1% 3% 1% 0% 0% 0% 15% 3% 2% 0% 8% 0% 3% 4% 2% -2% 2% 1% 5% -6% 0% -2% 43% 68% 4% 5% 0% 3% -5% 8% 8% 6% -10% -1% -2% -5% -9% 0% -7% 0% 9% 1% 1% 0% -2% -9% 3% 0% -1% GFDL A1Fi TPA Change year from current 2100 climateb 8.1 7.3 1.8 25.0 2.0 0.1 1.3 0.3 5.2 5.6 74.5 0.2 0.9 23.4 114.8 6.8 60.8 -43% 2% -10% -19% -47% 87% -46% 81% 45% 4% 11% 0% -17% -56% 10% -7% 5% Assessment area values were derived from the weighted average of secions. b Change represents the diference from current climate in 2100 and represents potenial change due to climate change. 123 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Several notable changes are predicted for current climate by the end of the century due to succession and management. Changes for current climate generally represent decreases in BA or TPA of short-lived and relatively shade intolerant species and increases in longer-lived, more shade tolerant, or faster growing species. For example, eastern redcedar, black oak, scarlet oak, black cherry, pignut hickory, and eastern white pine are projected to decline in BA and TPA whereas the more shadetolerant or longer lived white oak, sugar maple, and red maple are projected to increase (Figs. 38 and 39). The fast-growing and competitive species, such as tulip tree, increased more in TPA than BA, indicative of regeneration and growth after mortality. With a few exceptions, there were small to moderate differences in BA and TPA predicted for PCM B1 and GFDL A1FI compared to current climate by the end of the century (Table 22). Differences tended to be greater for GFDL A1FI than PCM B1, which is consistent with the Tree Atlas and LINKAGES results, and is attributed to the projections of higher average temperatures at the end of the century. The modest size of differences due to climate by the year 2100, especially given the potential for change indicated by the Tree Atlas and LINKAGES, is partly because trees are long-lived and turnover in species composition takes time. Species that showed declines across the region in BA (between 2 and 10 percent) and in TPA (between 17 and 5 percent) for GFDL A1FI by 2100 were American beech, Figure 39.—Projected changes in trees per acre for 17 species across the assessment area for the current climate scenario. Assessment area values were derived from the weighted average of secions. The width of the colored line represents trees per acre for each species at various points through ime. For example, red maple had the most trees per acre in 2010, but American beech is projected to become most abundant by the end of this century. 124 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS eastern hemlock, eastern white pine, scarlet oak, and sugar maple. The LINKAGES model predicted large decreases in potential growth for these species, for some to zero (extirpation). Species that generally increased in BA (up to 9 percent) and in TPA (up to 87 percent) for GFDL A1FI by the end of the century were loblolly pine, northern red oak, red maple, and tulip tree. The LINKAGES model also predicted large increases in potential growth for loblolly pine and tulip tree, and decreases or increases in northern red oak depending on section. Simulations by LANDIS PRO for 300 years into the future are not presented here because they are outside the scope of this assessment, but they show additional changes in species abundances in the directions suggested by the Tree Atlas and LINKAGES. Care should be used when interpreting values of percentage change because a large percentage change can occur for a small absolute change in BA or TPA if the initial values of BA or TPA were very small (Appendix 4). Geographic Trends The geographic and biological complexity of the assessment area warranted a closer look at the five ecological sections within the broader assessment area (see Appendix 4 for complete model results). For some species, BA or TPA are projected to increase in some sections while decreasing in others. For example, although northern red oak is projected to increase within the assessment area, these increases are largely concentrated in sections in northern Ohio (221F), and the easternmost sections (M221A and M221B) (Appendix 4). Although eastern hemlock is projected to decrease across the assessment area, most of the decrease in BA is projected in section 221E. Likewise, much of the decrease in basal area for chestnut oak is projected for GFDL A1FI in 221E (decrease of 14 percent), with no change projected in 221F and a 9-percent increase in M221A. Outputs from LANDIS PRO can be visualized spatially, and can provide greater context for interpreting the projected changes in tree volume. Figure 40 illustrates projected changes in basal area for northern red oak. It is important to note that these maps detect relative change on a pixel by pixel basis rather than averaged by section within the assessment area, as presented above. Figure 40.—Projected changes in basal area of northern red oak in 2100 compared to basal area in 2009. Basal area in 2100 under current climate represents the change in basal area atributable to succession and management. Basal area in 2100 for GFDL A1FI represents the total change in basal area atributable to succession and management plus high greenhouse gas emissions. 125 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS DiSCuSSioN oF MoDEL RESuLTS The three different models used in this assessment represent different facets of potential forest change as a result of a changing climate. Therefore, the ability to make comparisons between the different models facilitates a deeper understanding of which parts of a forest ecosystem may be most responsive or vulnerable to change. However, the differences between the models, in terms of design, outputs, strengths, and weaknesses, also make direct comparisons among model results difficult. This section describes areas of agreement and disagreement between the results and provides context for how the results from multiple models can be integrated to better understand forest change. Agreements The DISTRIB model used by the Tree Atlas allowed the most species to be modeled, and the LANDIS PRO model allowed the fewest. Therefore, only 17 species can be compared across all three models. The LANDIS PRO model demonstrates that forests are changing due to succession and management even without climate change. Succession and management will likely continue to be the most significant drivers of change over the next century, but actions that accelerate succession (e.g., management or natural disturbances) can facilitate climate-related changes. The LANDIS PRO model shows the beginnings of change in the directions suggested by LINKAGES and the Tree Atlas, but climate-related changes are too small to be conclusive for many of the species modeled. Where the Tree Atlas and LINKAGES are consistent with each other, results have the most certainty. All three models suggest that conditions for some species (e.g., American beech, eastern hemlock, eastern white pine, red spruce, and sugar maple) will become unfavorable by the end of the century for the higher climate scenario (GFDL A1FI). At the same time, all three models suggest that conditions 12 for other species (e.g., eastern redcedar and loblolly pine) will become more favorable by the end of the century, especially for GFDL A1FI. Additionally, both the Tree Atlas and LINKAGES tend to agree that many species will remain stable or increase for PCM B1 conditions and decrease for GFDL A1FI. These results support the idea that GFDL A1FI represents a future climate that is beyond the tolerance of many species. Additionally, these results suggest that many temperate species currently present in the assessment area could tolerate a mild degree of warming with corresponding increase in growing season precipitation, as represented by the PCM B1 scenario. Disagreements There do not appear to be any major discrepancies between individual species when LANDIS PRO and Tree Atlas results are compared, although there are some differences. The LANDIS PRO model projects changes (increase or decrease) in basal area of less than 20 percent for each species, but larger changes in trees per acre, particularly for GFDL A1FI, suggesting that young trees on the landscape will increase for several species. Whereas LANDIS PRO projects northern red oak to increase for GFDL A1FI, the DISTRIB and LINKAGES models project small decreases in suitable habitat and potential growth. This may be explained by LANDIS PRO’s ability to simulate changes in tree growth and biomass, whereas DISTRIB and LINKAGES describe potential suitable habitat that is available to a species. Although the amount of suitable habitat may decline, the remaining habitat may continue to be favorable for northern red oak, including the regeneration of northern red oak (in the absence of herbivory, competition, or other stressors). For many of the species above, LINKAGES and DISTRIB suggest great potential for landscape change in terms of basal area and trees per acre. However, LANDIS PRO results suggest that much of the change in forests in the next 100 years will ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS be due to succession, management, and disturbance; climate-related changes may take longer to manifest because trees are long-lived and disperse slowly. Changes in species establishment probability, and the consequences of losing habitat suitability, may not become evident by the end of the century, although changes in climate may already be setting the stage for substantial long-term changes in species composition that may include extirpation of some species and large expansion for others. Limitaions All models are simplified representations of reality, and no model can fully consider the entire range of ecosystem processes, stressors, interactions, and future changes to forest ecosystems. Each model omits processes or drivers that may critically influence ecosystem change in the future. Future uncertainty is not limited to climate scenarios; there is also uncertainty associated with future human interactions with forests. Examples of factors that are not considered in these models are: • Land management and policy responses to climate change or impacts to forests • Land-use change or forest fragmentation • Future changes in forest industry, including products and markets • Changes in phenology and potential timing mismatches for key ecosystem processes • Responses of understory vegetation, soil microorganisms, or soil mycorrhizal associations • Extreme weather events, which are not captured well in climate data or forest impact models • Future wildfire behavior, fire suppression, and ability to apply prescribed fire • Novel successional pathways for current forest ecosystems • Major insect pests or disease agents • Future herbivory pressure, particularly from white-tailed deer • Interactions among all these factors Most of these factors could drive large changes in forest ecosystems throughout the assessment area, depending on how much change occurs in the future. The potential for interactions among these factors adds layers of complexity and uncertainty. Despite these limitations, impact models are still the best tools available and can simulate a range of possible future outcomes. It is important to keep the above limitations in mind when weighing the results from different models and use them to inform an overall assessment. In the following section, we draw upon published literature to address other factors that may influence how forest ecosystems in the assessment area respond to climate change. SuMMARY oF CuRRENT SCIENTIFIC KNOWLEDGE oN FoREST iMPACTS The results presented above provide us with important projections of tree species distributions across a range of future climates, but these models do not account for all factors that may influence species and communities for a changing climate. Climate change has the potential to alter the distribution, abundance, and productivity of forests and their associated species in a variety of ways. These can broadly be divided into the direct effects of temperature and precipitation on forests and the indirect effects on forests through the alteration of current stressors or the development of additional stressors. For the most part, models such as the ones described above consider only direct effects such as average temperature and precipitation. Changes to forest management methods and their interaction with climate change may yield different outcomes. It is also important to note that some of the impacts may in fact be positive or beneficial to native forest ecosystems. The remainder of this chapter summarizes the current state of scientific knowledge on additional direct and indirect effects of climate change on forests in the assessment area and throughout the eastern United States. 127 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Drought Stress and Mortality There is evidence for an increased risk of future drought stress in the assessment area (see Chapter 4). Temperatures are expected to rise over the next century, and evapotranspiration in ecosystems is expected to increase as a result. Moisture stress may occur when increases in evapotranspiration are not offset by a corresponding increase in precipitation and soil moisture. Within the assessment area, the potential for more frequent droughts and moisture stress during the growing season appears to be much greater for the GFDL A1FI scenario (Chapter 4). However, for the milder PCM B1 scenario, warmer temperatures may also lead to increased evapotranspiration and physiological stress if increases in precipitation do not correspond to temperature increases. Additionally, there is evidence that precipitation is more likely to occur during larger precipitation events, which may increase the interval between rainfall events (Diffenbaugh et al. 2005). Drought can affect forests in many ways, including altering ecosystem processes, reducing forest productivity, increasing susceptibility to other stressors, and increasing tree mortality (Dale et al. 2001). Nearly all forests are susceptible to drought. For example, a recent study found that forests in both wet and dry environments around the world typically operate within a relatively narrow range of tolerance for drought conditions (Choat et al. 2012). Drought stress causes air bubbles to form in the xylem of growing trees (cavitation), which reduces the ability of trees to move water and causes reduced productivity or mortality. Forest species from rain forests, temperate forests, and dry woodlands all show a similarly low threshold for resisting droughtinduced cavitation (Choat et al. 2012). The potential effects of drought on forests will depend upon a number of factors, including drought duration and severity, as well as site-level characteristics of the forest. High stand density may 128 compound susceptibility to moisture stress because high-density stands face increased competition for available moisture (Keyser and Brown 2014, Olano and Palmer 2003). Additionally, drought-stressed trees are typically more vulnerable to insect pests and diseases (Dukes et al. 2009). Blowdowns Together with fire and ice, wind is a primary natural disturbance within the assessment area (Franklin et al. 2007, Ohio Department of Natural Resources [ODNR] 2010b). Blowdowns from large and small windstorms can have an important influence on the structure and species composition of forests (Abrams et al. 1995, Peterson 2000). Although tornadoes are relatively infrequent, intense winds generated from hurricanes, micro-bursts, and other storms can cause small patches of trees to uproot, especially on steep slopes (ODNR 2010b; Ulbrich et al. 2008, 2009). Hurricanes affecting the east coast can cause significant wind damage and blowdowns as far inland as western Maryland and West Virginia, where wind speeds can reach 50 miles per hour (Boucher et al. 2005). To date, the amount of evidence of changes in extreme storms in this region is rather limited (Dale et al. 2001, Intergovernmental Panel on Climate Change [IPCC] 2012). Some model projections suggest there may be an overall increase in the average wind speed in the area, but models disagree on whether trends in extreme cyclone frequency and intensity are increasing or decreasing (IPCC 2012, Ulbrich et al. 2009). If wind speeds do increase, the species that are most susceptible to blowdowns will likely differ by location across the assessment area. Blowdowns appear to disproportionately affect larger trees, shallow-rooted species, and thinned stands (Boucher et al. 2005, Dale et al. 2001). Sugar maple, sweet birch, and yellow birch are generally more wind resistant than black cherry, red maple, and tulip tree (Peterson et al. 2013). More frequent or widespread blowdown events can be expected to release the understory and accelerate the transition to ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS shade-tolerant species (Abrams and Scott 1989). This is especially the case in fire-dependent communities where shade-tolerant understories have developed in the absence of fire (Abrams and Nowacki 1992, Holzmueller et al. 2012). Gap-creating events may open up opportunities for regeneration of intermediate shade-tolerant species such as white oak, flowering dogwood, and various hickory species, especially at higher elevations (Abrams et al. 1998, Campbell et al. 2005). Blowdowns will continue to be an important ecological process in many Central Appalachians ecosystems, but existing scientific literature provides no clear indication of how blowdowns will be affected by the changing climate. create gaps, allowing regeneration of species such as red maple. If these events decrease or are eliminated from the area, new recruitment opportunities from this disturbance type may be limited. Winter Storm Damage Hydrology is tightly linked to the health and function of forest ecosystems, whether through maintenance of soil moisture during the growing season, seasonal flooding, creating necessary decomposition conditions, or other processes. Many forest systems in the assessment area have particular soil moisture requirements for the seasonality and extent of saturation. Additionally, certain species such as eastern cottonwood, eastern hemlock, and red spruce have particular seedbed requirements that are tightly linked to hydrologic conditions (Burns and Honkala 1990, Cornett et al. 2000). Snow and ice damage occurs occasionally across the area, and varies substantially with topography, elevation, exposure, and extent (ODNR 2010b). The most common cause of ice formation is when a winter warm front passes over much colder air. As rain falls from the warm layer through the layer at or below 32 °F, it becomes supercooled and able to freeze onto any surface it encounters (Turcotte et al. 2012). Although the number of days cold enough for snow and ice is projected to decrease, the intensity of these events when they do occur is projected to increase (Chapter 4). The decurrent growth habit (a wide crown with secondary trunks emerging from a main trunk) of many northern hardwoods makes them more vulnerable to ice damage than trees with a central leader (Turcotte et al. 2012). Species such as oaks, hickories, maples, and ashes appear to be particularly susceptible to branch and stem breakage, whereas conical species such as spruce are less susceptible. A study of a 2003 ice storm in Ohio found that oaks were more likely to show dieback than maples, and red maples were more likely to show dieback than sugar maples (Turcotte et al. 2012). Within species, damage appears to be greater in older, taller individuals, with higher mortality in sawtimber size classes than in pole or sapling size classes (Turcotte et al. 2012). These events also Although snow and ice will likely decrease across the area, some evidence suggests that storm events will actually increase during the winter months (Wang and Zhang 2008). With the projected increase in winter temperature, these events will more likely result in flooding and wind damage than in snow and ice damage, suggesting winter storms will function more like summer storms across the region. hydrologic impacts on Forests Climate change is likely to alter hydrologic regimes throughout the assessment area. As discussed in Chapters 3 and 4, heavy precipitation events have been increasing across the assessment area over the past century and this trend is expected to continue. In addition to more episodic precipitation events, future climate scenarios also project a wide possible range of seasonal precipitation and soil moisture (Chapter 4). Such variability may expose forest ecosystems to greater risk of hydrologic extremes: water-logging and flooding on one hand, and moisture stress and drought on the other. Forests that are accustomed to seasonal or annual variations in water availability may be better able to tolerate this variability. 129 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS A stream meandering through a small stream riparian forest. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. In a review of the consequences of precipitation variability on forests, Knapp et al. (2008) proposed that extreme precipitation events with longer intervals between events could have positive or negative impacts on a system, depending on its typical state in regards to soil moisture thresholds (Knapp et al. 2008). For example, xeric systems (adapted to dry conditions) would generally be less affected by dry periods because they are already limited by moisture stress, and larger precipitation events could recharge soil water levels, allowing for slightly longer periods of moisture. On the other end of the spectrum, hydric (i.e., wetland) systems are limited by anoxia rather than soil moisture, so longer dry periods between precipitation events would lower the water table, allowing oxygen to reach the roots of aquatic plants and ultimately increasing biomass productivity. Mesic systems (adapted to moderately moist conditions) would be the most affected by the increasing duration and severity of soil water stress because they are not well adapted to prolonged dry periods. This conceptual framework does not incorporate modiiers like soil texture and root depth, but the general principles are useful. 130 Flooding can affect forest systems differently, depending on the frequency and duration of loods, and the soil, vegetation, and topographic complexity of the landscape. In mountainous areas, loods are generally brief and intense, with loodwaters funneling rapidly down steep slopes and into valley streams (Eisenbies et al. 2007, Swanson et al. 1998). These swift, ierce loods often damage trees by breaking stems and limbs, and scouring vegetation. In lowland areas, loods are generally more gradual and last longer, with longer periods of soil saturation and less tree breakage. Extreme or very heavy precipitation events can also have important consequences on riparian and lowland systems when they result in looding, which can increase erosion and transport of nutrients, contaminants, and pathogens (Groffman et al. 2014). Disturbances caused by loods, drought, scouring by ice, and river channeling often drive tree species and forest diversity, especially in lowland and riparian forests (Vadas and Sanger 1997). ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS Soil Erosion Wildire As climate change continues to intensify the hydrologic cycle, the increase in heavy rainfall events is projected to continue across the assessment area. One of the potential impacts of this trend is that soil erosion rates will increase (Nearing et al. 2004). Soil erosion is considered one of the major threats to the Central Appalachians region, but many studies examining the effects of climate change on soil erosion have focused on agricultural settings, rather than forest ecosystems (Pennsylvania Department of Environmental Protection [PDEP] 2012). Although additional vegetative cover and root stabilization in forest systems may make forests less prone to soil erosion, not all forest soils will be equally protected. Reductions in vegetative cover due to a variety of climate-related impacts, such as prolonged drought, could lead to an increase in susceptibility to erosion. Additionally, the intensification of precipitation changes combined with orographic effects may increase risk or severity of erosion in mountainous areas (Beniston 2003, Sturdevant-Rees et al. 2001). Wildfire was historically an important driver for some forest ecosystems in the assessment area, although it has been largely suppressed since the 1930s. In contrast to the large wildfires that occur periodically in the western United States, contemporary wildfire events in the eastern United States consist of numerous small fires in the wildland-urban interface (Peters et al. 2013). Development and fragmentation in the form of high housing density are the biggest source of ignition, but access to the surrounding infrastructure allows wildfires to be extinguished relatively quickly (Bar Massada et al. 2009). Ignitions are caused primarily by humans and less frequently by lightning (National Interagency Fire Center 2013). The conditions responsible for wildfire behavior are the result of weather, topography, and fuels (Moritz et al. 2012). Climate can directly affect the frequency, size, and severity of fires, and indirectly affect fire regimes through effects on vegetation vigor, structure, and composition (Moritz et al. 2012, Sommers et al. 2011). Soil erosion is also closely correlated with precipitation. One study estimates that for every 1-percent increase in rainfall, runoff could increase by 2 percent, and erosion could increase by 1.7 percent (Nearing et al. 2004). Another study examined changes in erosivity across the United States at a very large grid scale and found that erosion may increase or decrease in the assessment area depending on the climate model used (Nearing 2001). This study looked only at broad-scale changes in precipitation, and does not account for other factors that may affect the vulnerability of soil to erosion such as vegetation cover, slope, or soil type. Reductions in biomass and vegetative cover due to climate change impacts or land-use changes (e.g., forest roads) could also lead to an increase in erosion susceptibility (Nearing 2001). Invasive species may also interact with climate to increase the frequency, intensity, or length of the fire season (Brooks and Lusk 2008). Invasive shrubs and herbs may increase the density of the understory, thereby increasing fuel. On the other hand, many invasive shrubs and herbs begin growing earlier in spring than native plants. This early green-up may reduce the flammability of fire-adapted communities during the spring fire season (Brooks and Lusk 2008). Invasive pests can also interact with climate and wildfire by altering forest fuels and forest structure (Ehrenfeld 2010, Krist et al. 2007, Lovett et al. 200, Szlavecz et al. 2010). 131 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS An analysis of fire probability across the globe using 1 downscaled climate models found low agreement among projections of climate change effects on fire probability in the central United States in the near term (2010 to 2039), but most models projected an increase in wildfire probability by the end of the century (2070 to 2099) (Moritz et al. 2012). This agreement is particularly high for temperate coniferous forests and temperate broadleaf and mixed forests, where fire probability models were most sensitive to mean temperature of the warmest month. If temperature and evapotranspiration increase drying of the forest floor in spring, amplify the effects of declining precipitation, or overwhelm modest precipitation increases, the annual area burned and length of the fire season will likely increase. Projected increases in lightning-producing convective storms may also increase ignition frequency (Sommers et al. 2011). Another global study using a sensitive model and a high emissions scenario projected increased fire potential across the United States, including the assessment area (Liu et al. 2010). Duration of the fire season is projected to lengthen by several months by the end of the century, primarily due to warming (Liu et al. 2010). How a change in fire risk across the region translates to effects at local scales in forests of the assessment area also depends on land use and management decisions. Fire suppression has already been linked with mesophication in eastern forests, and fire management is expected to continue to drive vegetation and succession in the future (Nowacki and Abrams 2008). To understand how climate change may interact with wildfire in the United States, model simulations of vegetation cover types were conducted for high and low emissions scenarios (A2 and B2; see Chapter 2) with wildfire suppressed and unsuppressed for the period 2070 through 2099 (Lenihan et al. 2008). Under both suppressed and unsuppressed wildfire, the range of temperate deciduous forest across the eastern United States was projected to shift northward, with a 132 large loss of cool mixed forest. Under unsuppressed wildfire, some forest just outside the assessment area in Ohio was projected to transition to a woodland or savanna type, and there is potential that existing woodlands and savannas may expand where they do occur (Lenihan et al. 2008). Fire suppression does not allow this expansion; cool mixed forest is projected to be largely replaced by temperate deciduous forest as both biomes shift northward. Many aspects of the fire regime within the assessment area will likely be affected by changes in climate, with response to climate change varying over time and space. Dry-mesic oak, dry oak and pine-oak, and dry calcareous forests are often tied to wildfire dynamics, but fire could also become an increasing source of disturbance in other forest types if climatic shifts over the 21st century result in different fire behavior. Forest ecosystems adapted to dry habitat conditions (e.g., oak, pine, and hickory) may be the most likely areas to burn. Forest systems adapted to habitats with abundant moisture (e.g., northern hardwoods), and those reliant on ground seepage at higher elevations may be able to better compensate for increased evapotranspiration and higher temperatures. However, even these systems may be more likely to burn if projected temperature increases result in drier habitat conditions. Fire effects on nutrient availability depend not only on species composition but also on the intensity and duration of the fires (Certini 2005). Lowintensity fire can release nutrients, but higher fire temperatures may result in mineralization and volatilization, especially on acidic soils, which dominate most of the higher elevation portion of the assessment area (Gray and Dighton 200). A watershed-scale study of prescribed fire in southeastern Ohio found that periodic low-intensity prescribed fire can return nutrient cycling and microbial activity to presettlement levels, which can restore ecosystem functions of mixed oak forests (Boerner 200). Authors of a review paper on climate and wildfire conclude that fire-related ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS impacts may be more important to some ecosystems than the direct effects of climate change on species fitness and migration (Sommers et al. 2011). Fire could have a greater influence because it can be a catalyst for change in vegetation, perhaps prompting more rapid change than would be expected based only on the changes in temperature and moisture availability. As with wind disturbances, the potential exists for novel successional pathways after wildfire if climatic conditions, seed sources, or management decisions favor different forest types. increases in Carbon Dioxide In addition to effects on climate, carbon dioxide itself can affect plant productivity and species composition. Elevated carbon dioxide may enhance growth and water use efficiency of some species (Ainsworth and Rogers 2007, Norby et al. 2005), potentially offsetting the effects of drier growing seasons. This is commonly called “carbon dioxide fertilization.” There is already some evidence for increased forest growth in the eastern United States (Cole et al. 2010, McMahon et al. 2010, Pan et al. 2009), but it remains unclear if long-term enhanced growth can be sustained (Bonan 2008, Foster et al. 2010). Nutrient and water availability, ozone pollution, and tree age and size all play major roles in the ability of trees to capitalize on carbon dioxide fertilization (Ainsworth and Long 2005). Ecosystem community shifts may take place as some species are genetically better able to take advantage of carbon dioxide fertilization than others (Souza et al. 2010). Some models account for changes in carbon dioxide, but these models tend to focus on nutrient cycling and general vegetation types, and not specific species (Lenihan et al. 2008, Ollinger et al. 2008). Therefore, this assessment is not able to combine the effects of carbon dioxide fertilization with the effects of temperature and precipitation on particular species. Changes in Nutrient Cycling and Acid Deposiion As air temperatures warm and precipitation patterns change, the way nutrients are cycled between plants, soils, and the atmosphere may also change. The long-term effects of acid deposition have an added effect that makes this cycle more complex and hard to predict in the future. Increases in droughts and floods, changes in phenology, and the interaction among these factors can also impair nutrient cycling and the availability of nitrogen to trees and other vegetation (Rennenberg et al. 2009). Decomposition of vegetation is carried out primarily by enzymes released from bacteria and fungi. These enzymes are sensitive to changes in temperature, and thus there is generally a positive effect of temperature on the rate of enzymatic activity as long as moisture is also sufficient (Brzostek et al. 2012, Finzi et al. 200, Rustad et al. 2001). A number of studies have examined the effects of extended dry periods followed by moisture pulses on nutrient cycling (Borken and Matzner 2009). Although these moisture pulses do lead to a flush of mineral nitrogen, it is not sufficient to compensate for the lack of microbial activity during dry periods. Thus, an increase in wet-dry cycles appears to lead to a reduction in nutrient availability for trees. Although warmer temperatures have the potential to increase enzymatic activity and nutrient cycling, acidification will remain an important consideration. Anthropogenic emissions of nitrogen and sulfur have increased over the last century, peaking in the 1970s. These emissions undergo chemical transformations that produce nitrates and sulfates, which are eventually deposited on the ground (Elliott et al. 2013). These sulfur and nitrogen compounds are also deposited at high concentrations through rain and snow in the eastern United States, particularly 133 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS in high-elevation sites (Pardo et al. 2011). In forest ecosystems, hydrogen ions associated with nitrogen and sulfur deposition replace nutrient base cations of calcium, magnesium, and potassium, depleting these nutrients and allowing them to leach into drainage waters. At the same time, toxic cations of aluminum are mobilized, and the combined effects of nutrient depletion and increased toxicity have been proven to reduce the health and productivity of forests and streams through acidification (Elliott et al. 2013, Long et al. 2013, Schaberg et al. 200). Nitrogen saturation has also been shown to reduce carbon allocation to plant roots and mycorrhizae (Pardo et al. 2011). Acid deposition has likely contributed to the increased susceptibility of forests to drought and insect attack, and is expected to contribute to reduced ability to withstand climate changes (Friedland et al. 1984, McNulty and Boggs 2010, Pardo et al. 2011). Researchers simulating the effects of nitrogen and sulfur deposition on wilderness areas in North Carolina found that even with dramatic reductions in acid deposition, ecosystems will take decades to recover from the effects of acidification (Elliott et al. 2013). Future levels of nitrogen and sulfur deposition can be controlled through efforts to significantly reduce air pollution by 204 (e.g., through the Clean Air Act amendments of 1977). In the interim, projected increases in winter and spring precipitation could facilitate the deposition of air pollutants. The effects of climate change on nutrient cycling will likely be overshadowed by the impacts of nitrogen and sulfur deposition in the assessment area over the next 50 years or longer. space, water, nutrients, and light (Brown and Peet 2003, Dukes et al. 2009). Wetland and riparian areas are particularly susceptible to nonnative plant invasion, partially due to passive seed dispersal via surface waters (Nilsson et al. 2010). Invasive species in riparian areas are likely better competitors for nutrient pulses supplied by runoff (PDEP 2004). Some of the most prolific riparian invasives are the mile-a-minute vine, purple loosestrife, Japanese knotweed, common reed, Japanese stiltgrass, and reed canarygrass. Many invasive species that currently threaten forests in the Central Appalachians region may benefit directly from projected climate change or benefit from the slow response of native species (Rebbeck 2012). Increases in carbon dioxide have been shown to have positive effects on growth for many plant species, including some of the most invasive weeds in the United States (Ziska 2003). Experiments with carbon dioxide fertilization on kudzu seedlings have indicated increased growth, increased competition with native species, and range expansion (Sasek and Strain 1988, 1989). Models have also projected that increased carbon dioxide emissions and subsequent warmer winter temperatures will likely expand the Invasive Plant Species As described in Chapter 1, nonnative invasive species are a major threat to all forest ecosystems across the eastern United States. Many invasives are able to establish rapidly following a disturbance, and are able to outcompete native vegetation for growing 134 Naive grasses and forbs, dominant plants in this lat, wet area. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS northern ranges of bush honeysuckle (Sasek and Strain 1990), privet, kudzu, and cogongrass (Bradley et al. 2010). Cogongrass in the southeastern United States has contributed to altered fire regimes and is expected to advance northward with warmer temperatures (Lippincott 2000). Some invasive species are tolerant of drought and fire, and may be at an even greater advantage for future climate conditions. Ailanthus and bush (amur) honeysuckle are woody invasives that currently have negative impacts on forests across the assessment area (Hutchinson and Vankat 1998). In addition to directly outcompeting native oak-hickory trees by rapid growth, ailanthus and bush honeysuckle have allelopathic effects on soils, exuding a toxin that discourages the growth of other plants (ODNR 2011b, Williams 2005). Other species, such as garlic mustard and Japanese stiltgrass, are not particularly drought-tolerant, but their persistent seed banks enable them to recover in wetter years (Fryer 2011, Munger 2001). Milder winters may allow some invasive plant species to survive farther north than they had previously (Dukes et al. 2009). For example, kudzu is a drought-tolerant invasive vine that has invaded forests in the southeastern United States and has already appeared within the assessment area (Munger 2002). The northern distribution of kudzu is limited by cold winter temperatures, and models have projected that warmer temperatures will result in expansion northward (Bradley et al. 2010). Chinese and European privet are invasive flowering shrubs that crowd out native species and form dense thickets. Habitat models project increased risk for privet invasion into Ohio, West Virginia, and Maryland by the end of the century (Bradley et al. 2010). insect Pests and Pathogens Warmer temperatures, moisture deficit, and compounding stressors may increase the susceptibility of trees to insect pests and pathogens (Weed et al. 2013). Warmer winter temperatures may also result in increased abundance of pests and pathogens that are currently present in the assessment area. For example, hemlock woolly adelgid populations are currently limited by low winter temperatures and freeze-thaw cycles, and populations of hemlock woolly adelgid have increased or expanded northward during mild winters (Pennsylvania Department of Conservation and Natural Resources 2013). The emerald ash borer, currently devastating populations of ash species, has been observed to produce more generations under warmer conditions (DeSantis et al. 2013, Venette and Abrahamson 2010, Wei et al. 2007). Other pest outbreaks, including those of native species (e.g., forest tent caterpillar and spruce budworm), are more common when trees are stressed by factors such as drought (Babin-Fenske and Anand 2011, Gray 2008). The interacting effects of drought and increased pests and pathogens may result in increased risk of oak decline, which is largely driven by insect pests and pathogens predisposed to invasion in drought conditions (Clatterbuck and Kauffman 200, McConnell and Balci 2014). There is evidence that other species may be disadvantaged by climate change; for example, the hatching of gypsy moth eggs is dependent on the budburst of host trees. Changes in phenology could result in starvation if the eggs hatch before budburst (Ward and Masters 2007). Tree pathogens, such as the fungus Armillaria mellea, can also potentially increase in abundance and range, and may result in increased disease that stresses or kills forest trees. Armillaria populations will likely increase with increasing temperatures, and become a more severe threat during drought periods, when host trees are more susceptible to root diseases (Kliejunas 2011). Warmer temperatures will also increase the susceptibility of tree species to pests and diseases that are not currently a problem in the assessment area (Logan et al. 2003). Projections of gypsy 135 ChAPTER 5: FuTuRE CLiMATE ChANGE iMPACTS oN FoRESTS moth population dynamics for a changing climate suggest substantial increases in the probability of establishment in the coming decades (Logan et al. 2003). Oak species that would otherwise do well in a changing climate could consequently be at risk. In addition, future northward range expansion attributed to warming temperatures has been projected for southern pine beetle (Ungerer et al. 1999). A recent outbreak of southern pine beetle in New Jersey has already been attributed to warmer temperatures (Weed et al. 2013). Southern pine beetle could become a threat if shortleaf pine expands in the region. Efects of Vertebrate Species Herbivory, seed predation, and disturbance by vertebrates can be important stressors in the Central Appalachians region. Currently, little is known about how these factors could be affected by climate change. Deer overbrowsing and seed predation may reduce the overall success of species that are otherwise projected to do well under future climate change (Ibáñez et al. 2008). For example, white oak is projected to increase in habitat suitability and basal area, but the models mentioned earlier in this chapter do not account for the herbivory of young oak regeneration by deer. Currently, there is little evidence to indicate how deer and other vertebrate species will respond to climate change in the assessment area. An analysis of climate change impacts on white-tailed deer in Wisconsin suggests that deer in that area will likely be subject to a mixture of positive impacts from milder winters coupled with negative impacts from increased disease outbreaks (Wisconsin Initiative on Climate Change Impacts 2011). How these two factors may influence deer populations in Ohio, West Virginia, and Maryland remains unknown. 13 ChAPTER SuMMARY Although models are useful for exploring potential future changes, all models are simplified representations of reality, and no model can fully consider the entire range of ecosystem processes, stressors, interactions, and future changes to forest ecosystems. The DISTRIB (Tree Atlas), LINKAGES, and LANDIS PRO models suggest that conditions for some species (e.g., American beech, eastern hemlock, eastern white pine, red spruce, and sugar maple) will become unfavorable by the end of the century for GFDL A1FI. At the same time, all three models suggest that conditions for other species (e.g., eastern redcedar and loblolly pine) will become more favorable by the end of the century, especially for GFDL A1FI. Additionally, the Tree Atlas and LINKAGES tend to agree that many species will remain stable or increase for PCM B1 and decrease for GFDL A1FI. These results support the idea that GFDL A1FI future climate is beyond the tolerance of many species, but that many species could tolerate a mild degree of warming with a corresponding increase in growing season precipitation, as represented by PCM B1. Generally, the changing climate tends to intensify the stressors that may already exist for many species and increases susceptibility to drought, pests, diseases, or competition from other species. It is the interaction among all these factors that will drive the response of forests to climate change. All of these factors need to be taken into account when evaluating the vulnerability of Central Appalachians forests to climate change. The vulnerability of nine forest ecosystems is described in Chapter . ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Changes in species distribution and abundance due to climate change can have important implications for the habitats in which those species live, leading to shifts in community composition and changes in ecosystem processes (Janetos et al. 2008, Vose et al. 2012). In addition, climate change itself can alter ecosystem drivers and exacerbate or ameliorate current stressors (Janetos et al. 2008, Vose et al. 2012). This chapter describes the climate change vulnerability of nine major forest ecosystems in the Central Appalachians assessment area (see Chapter 1 for a description of the nine forest ecosystems). Vulnerability is the susceptibility of an ecosystem to the adverse effects of climate change (Intergovernmental Panel on Climate Change [IPCC] 2007). It is a function of the potential impacts (a combination of exposure and sensitivity) to an ecosystem and the adaptive capacity of the ecosystem to tolerate those impacts (Fig. 41). We consider a forest ecosystem to be vulnerable if it is at risk for no longer being recognizable as that ecosystem, or if the ecosystem is anticipated to suffer substantial declines in health or productivity. We considered the vulnerability of an ecosystem to climate change independent of the economic or social values associated with the ecosystem, even though forest management, land-use changes, and human population pressures can have dramatic and immediate effects on ecosystems. The ultimate decision of whether to use resources to try to conserve a vulnerable ecosystem or allow it to shift to an alternate state will depend on the individual objectives and resources of land management organizations. This chapter is organized into two sections. First, we present an overall synthesis of potential climate impacts on forest ecosystems, organized according to drivers and stressors, ecosystem impacts, and factors that influence adaptive capacity. This synthesis is based on the current scientific consensus of published literature (Chapters 4 and 5). In the second section, we present individual vulnerability determinations for the nine forest ecosystems considered in this assessment. Figure 41.—Key components of vulnerability, illustraing the relaionship among exposure, sensiivity, and adapive capacity. Adapted from Glick et al. (2011). 137 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES A SYNThESiS oF CLiMATE ChANGE iMPACTS oN FoREST ECoSYSTEMS Potential impacts are the direct and indirect consequences of climate change on individual ecosystems. Impacts are a function of an ecosystem’s exposure to climate change and its sensitivity to those changes. Impacts could be beneficial to an ecosystem if the changes result in improved health or productivity, occupation of an expanded area, or a tendency to maintain the current identity of the ecosystem. Negative potential impacts would include declining health and productivity, reduced area occupied, or a composition shift that leads to a substantially different identity of the ecosystem. Throughout this chapter, statements about potential impacts and adaptive capacity factors are qualified with a confidence statement. These confidence statements are formatted according to a confidence determination diagram from the IPCC’s recent guidance for authors (Mastrandrea et al. 2010) (Fig. 42). Confidence was determined by gauging both the amount of available evidence and the level of agreement among that evidence. Evidence was robust when multiple lines of evidence were available in addition to an established theoretical understanding to support the vulnerability determination. Agreement refers to the agreement among the available sources of evidence, not among authors of this assessment. If theories, observations, and models tended to suggest similar outcomes, the sum of the evidence resulted in a high level of agreement. Potenial Impacts of Climate Change on Drivers and Stressors Many physical and biological factors contribute to the current state of forest ecosystems in the Central 138 Figure 42.—Conidence determinaion diagram used in the assessment. Adapted from Mastrandrea et al. (2010). Appalachians region. Some of these factors serve as drivers, or defining features that determine the identity of an ecosystem. Other factors can serve as stressors, reducing the health, productivity, and integrity of specific ecosystems. Many factors, such as flooding or fire, may be drivers in one ecosystem and stressors in another. Temperatures will increase (robust evidence, high agreement). All downscaled climate models project that average temperatures will increase across much of the assessment area. A large amount of evidence from across the globe shows that average temperatures have been increasing and will continue to increase due to human activities (Chapter 2). Temperatures across the assessment area have already been changing over the last century (Chapter 3), and temperature increases are projected even under the most conservative climate scenario, with dramatic increases projected under the high climate scenario (Chapter 4). ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Growing seasons will get longer (robust evidence, high agreement). There is high agreement among evidence that projected temperature increases will continue the current trend of longer growing seasons in the assessment area. Evidence at both global and regional scales indicates that growing seasons have been getting longer, and this trend is likely to become even more pronounced over the next century (Chapters 3 and 4). Longer growing seasons have the potential to affect the timing and duration of ecosystem and physiological processes across the region (Dragoni and Rahman 2012). As seasons shift so that spring arrives earlier and fall extends later into the year, plant species may respond to changes in temperature regimes with changes in the timing of leaf-out, reproductive maturation, and other developmental processes (Schwartz et al. 200a, Walther et al. 2002), and some of these changes have already been observed (McEwan et al. 2011). Longer growing seasons, especially those that are extended in the fall, could also result in greater growth and productivity of trees and other vegetation, but only if balanced by available water and nutrients (Chapter 5). Longer growing seasons could also lead to changes in the distributions of plant and animal species (Iverson et al. 2008b). The amount and timing of precipitation will change (medium evidence, high agreement). All downscaled climate models agree that there will be changes in precipitation patterns across the assessment area. For the climate projections used in this assessment (Chapter 4) and other publications, projected changes in precipitation are highly variable (Kunkel et al. 2013a, 2013b). The PCM B1 scenario projects annual precipitation to increase over most of the assessment area, but decrease sharply in the Allegheny Mountains section. The GFDL A1FI scenario projects increases over much of the Ohio and Maryland portions of the assessment area, and widespread decreases over larger areas of West Virginia, including the Allegheny Mountains (Chapter 4). Models also project changes in precipitation patterns between seasons (Kunkel et al. 2013b). Precipitation increases are expected under both scenarios in winter and spring, with larger increases under GFDL. Summer and fall precipitation projections suggest a wide range of potential responses, from decreases to increases, depending on the climate scenario (Chapter 4). Intense precipitation events will continue to become more frequent (medium evidence, medium agreement). There is some agreement that the number of heavy precipitation events will continue to increase in the assessment area. If so, impacts from flooding and soil erosion may also become more damaging. Total precipitation in the assessment area has increased the most in fall, by 2.3 inches over the last century. The timing and magnitude of precipitation events have shifted, however, so that more rain is falling during larger events, and light rainfall events are becoming less common (Chapter 3). Rainfall from these high-intensity events represents a larger proportion of the total rainfall, meaning that the precipitation regime is becoming more episodic with potentially longer dry spells between events. Climate models project this trend to continue through the end of the century, with an additional increase in the number of heavy precipitation events throughout the central and northeastern United States (IPCC 2012). Ecosystems are not all equally capable of holding moisture that comes in the form of extreme events. Increases in runoff after heavy precipitation events could also lead to an increase in soil erosion, specifically channel erosion (Nearing et al. 2004). 139 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Severe storms will increase in frequency and severity (medium evidence, medium agreement). There is some agreement that future climate change will destabilize atmospheric circulation patterns and processes, leading to increased risk of severe weather. Although the positive trend in historic storm frequency is muddled with greater public awareness, reporting, and recent advances in technology, future trends can be predicted by using atmospheric models. Projected increases in temperature, precipitation, and convective available potential energy over the next century are expected to result in more frequent days when conditions are favorable for severe storms (Trapp et al. 2007, 2009, 2011). Many storms affecting the assessment area are generated in the southwestern United States and from the Atlantic Ocean; therefore changes in conditions in these regions may contribute to increased frequency and severity of storms within the assessment area (Trapp et al. 2007). Soil moisture patterns will change (medium evidence, high agreement), with drier soil conditions later in the growing season (medium evidence, medium agreement). Studies show that climate change will have impacts on soil moisture, but there is some disagreement among climate and impact models on how soil moisture will change during the growing season. As discussed above, seasonal changes in precipitation are expected across the assessment area. Due to potential decreases in summer and fall precipitation and increases in winter and spring precipitation, it is likely that soil moisture regimes will also shift. Longer growing seasons and warmer temperatures may also result in greater evapotranspiration losses and lower soil-water availability later in the growing season (Chapter 4). The Variable Infiltration Capacity model projected 140 summer and fall decreases in soil moisture, with the greatest decrease (10 percent) in the West Virginia portion of the assessment area (Ashfaq et al. 2010). How these broad trends affect the Central Appalachians region will depend greatly on landscape and topographic position and therefore exposure to climate changes. South-facing slopes may be particularly vulnerable to soil drying in late summer or fall. Seedlings will be more vulnerable to these effects than mature individuals; just one severely dry growing season per decade may greatly reduce regeneration success of most species (Gómez-Aparicio et al. 2008). Climate conditions will increase wildfire risk by the end of the century (medium evidence, medium agreement). Some national and global studies suggest that wildfire risk will increase in the region, but few studies have specifically looked at wildfire potential in the assessment area. Although there is greater uncertainty around future fire behavior for the near term, model simulations tend to agree that there will be global increases in fire activity by the end of the 21st century (Moritz et al. 2012). The duration of the fire season in the Central Appalachians is closely linked with increases in average temperature during the summer (Liu et al. 2010). Interactions between complex patterns of land use and ownership, forest fragmentation, and both human and natural ignition sources, make it difficult to determine how an increase in fire weather conditions might be manifested. In addition to the direct effects of temperature and precipitation, increases in fuel loads from pest-induced mortality, exotic species invasion, or blowdown events could also increase fire risk (Lovett et al. 200, Weed et al. 2013). Forest fragmentation and unknown future wildfire management decisions may limit individual fires even as fire risk increases. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Certain insect pests and pathogens will increase in occurrence or become more damaging (medium evidence, high agreement). Evidence indicates that an increase in temperature will lead to increases in pest and pathogen outbreaks, but research to date has examined few species in the assessment area. Changes in climate may allow some pests and pathogens to expand their ranges, or to become a larger threat. Changes in climate may also increase tree species’ susceptibility to the entire suite of native and nonnative pests and pathogens, including hemlock woolly adelgid, southern pine beetle, and forest tent caterpillar. Pests and pathogens are generally more damaging in drought-stressed ecosystems, so there is high potential for these agents to interact with other climate-mediated stressors. For example, susceptibility of trees to sudden oak death is linked to periods of drought stress. The fungus Phytophthora ramorum, a known contributor to sudden oak death in California and Europe, is currently spread through nurseries and has appeared in nursery samples in Connecticut. The climate of the assessment area is favorable to this fungus, which is likely to increase in abundance and extent in association with wetter springs (Kliejunas 2011). Furthermore, the abundance of potential host species in the assessment area increases the threat from introduction of this new disease (Ockels et al. 2004). Unfortunately, we lack basic information on the climatic thresholds that apply to many forest pests, and our ability to predict the mechanisms of infection, dispersal, and transmission for disease agents remains low. It is also not possible to predict all future nonnative species, pests, or pathogens that may enter the assessment area during the 21st century. Many invasive plants will increase in extent or abundance (medium evidence, high agreement). Evidence indicates that an increase in temperature and more frequent disturbances will lead to increases in many invasive plant species. Many invasive species that currently threaten forests in the Central Appalachians region may benefit directly from projected climate change or benefit from the slow response of native species (Rebbeck 2012). Increases in carbon dioxide have been shown to have positive effects on growth for many plant species, including some of the most invasive weeds in the United States (Ziska 2003). Changes in climate may allow some invasive plant species to expand their ranges northward, such as bush honeysuckle, privet, kudzu, and cogongrass. Milder winters may allow some invasive plant species to survive farther north than they had previously (Dukes et al. 2009). Some invasive species are tolerant of drought and fire, and may be at an even greater advantage for future climate conditions. Future increases in fire or flooding are likely to benefit the many invasive plants that are able to establish quickly and outcompete native vegetation on disturbed sites (Brown and Peet 2003, Dukes et al. 2009). Increases in riparian flooding is expected to contribute to more frequent disturbance, and therefore higher impacts from invasive species. Potenial Impacts of Climate Change on Forests Shifts in drivers and stressors mentioned above will naturally lead to changes in forest ecosystems throughout the assessment area. Indirect impacts of climate change may be indicated by shifts in suitable habitat, species composition, or function of forest ecosystems. 141 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Suitable habitat for northern species will decline (medium evidence, high agreement). All three impact models project a decrease in suitability for northern species such as eastern hemlock, red spruce, and sugar maple, compared to current climate conditions. Across northern latitudes, past periods of warmer temperatures have resulted in changes in species distribution northward and to higher elevations (Chen et al. 2011, Parmesan and Yohe 2003). The ranges of eastern hemlock and red spruce are largely to the north of the assessment area, but these species currently persist in microhabitats that remain cool and moist enough to support them. Red spruce is more limited within the assessment area, occurring at high elevations in the Allegheny Mountains section of West Virginia. Hemlock is more widespread, occupying cool and wet sites at lower elevations. As these species’ ranges continue to shift northward, they may become rare or extirpated from the area. In the absence of other mortality agents, long-lived individuals already established in cool, wet microhabitats may persist for many years, even when habitat becomes unsuitable for regeneration or growth (Iverson and Prasad 1998). Due to the geographic limitations of these species’ current habitat, however, they are unlikely to migrate even if newly suitable habitat becomes available elsewhere in the assessment area. Results from climate impact models also suggest declines in suitable habitat Eastern hemlock. This species is threatened by the hemlock woolly adelgid. Photo by Patricia Butler, Northern Insitute of Applied Climate Science (NIACS) and Michigan Tech, used with permission. 142 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES for northern species that are not as geographically limited, such as sugar maple (Chapter 5). These species near the southern edge of their range may also be able to persist in southern refugia if potential new competitors from farther south are unable to colonize these areas, although they are expected to have reduced vigor and be under greater stress (Iverson et al. 2008b). Habitat is projected to become more suitable for southern species (medium evidence, high agreement). All three impact models project an increase in suitability for southern species such as eastern redcedar and loblolly pine, compared to current climate conditions. Model results project that tree species currently at their northern range limits south of the assessment area will become more abundant and more widespread under a range of climate futures (Chapter 5). The range of eastern redcedar is widespread throughout the eastern United States, but currently occupies a small portion of its range within the assessment area. The range of loblolly pine lies entirely south of the assessment area, although disjunct populations have been planted in some locations within Ohio and Maryland. Models agree that loblolly pine will fare well in terms of habitat and basal area, especially under GFDL A1FI. Post oak and shortleaf pine are also projected to fare well under both scenarios. The ranges of both species are largely south and west of the assessment area, with populations most abundant to the west. Blackjack oak, common persimmon, osage-orange, southern red oak, sugarberry, sweetgum, and winged elm are also projected to increase, but were modeled only by the Tree Atlas. Several species that do not currently exist within the assessment area are projected to have new suitable habitat: water oak, water locust, and cedar elm. However, habitat fragmentation and the limited dispersal ability of seeds are expected to hinder movement of these southern species despite the increase in habitat suitability (Ibáñez et al. 2008). Most tree species can be expected to migrate more slowly than their habitats will shift (Davis and Shaw 2001). Indeed, in a simulation for five tree species, a maximum of 15 percent of newly suitable habitat would have a chance of getting colonized over 100 years (Iverson et al. 2004a, 2004b). Pests and diseases such as fusiform rust, annosus root rot, and southern pine beetle may also limit the expansion of loblolly pine. As suitable habitat increases for some tree species and decreases for others, there will be new opportunities for species to become new components of novel forest types or commercial plantations (Iverson et al. 2008a). Species composition will change across the landscape (limited evidence, high agreement). Although few models have specifically examined how species composition may change, model results from individual species, paleoecological data, and ecological principles suggest that recognized communities may dissolve to form new mixes of species. Decoupling of the drivers, stressors, and dominant tree species that define forest ecosystems is expected to lead to a rearrangement of suitable conditions for tree species within the assessment area. This rearrangement may result in the dissolution of current plant community relationships, which paleoecological evidence shows occurred in the past (Davis et al. 2005, Root et al. 2003, Webb and Bartlein 1992). Canopy and understory species composition is closely tied to soil moisture, aspect, slope position, and other environmental variables (Hix and Pearcy 1997). Shifts in overstory and understory structure may follow somewhat predictable pathways based on shifts in soil moisture, fire frequency, and disturbance regime, but will still be strongly correlated to landscape position (Iverson et al. 1997). For example, on the Wayne National Forest, dominant tree species are expected to be oaks on dry ridge tops, and tulip tree and black cherry on mesic sites (Iverson et al. 1997). The model results presented in Chapter 5 raise the possibility for potentially large differences in 143 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES species’ responses across the Central Appalachians. Generally, the models indicate that climate trends may favor oaks and pines, although ecological lag times and management decisions may slow conversions of forest types. Repeated harvesting, grazing, and other large-scale disturbances have already created atypical relationships among the canopy and understory species in many areas, lending more uncertainty to future community composition (Root et al. 2003). If associated species such as pollinators and mycorrhizae do not migrate into newly suitable areas, further constraints could be placed on native species colonization (Clark et al. 1998). Furthermore, nonnative invasive plants may be better able to fill newly created niches (Hellmann et al. 2008). A major transition in forest composition is not expected until after the middle of the century (2040 to 2069) (medium evidence, medium agreement). Although some models indicate major changes in habitat suitability, results from spatially dynamic forest landscape models indicate that a major shift in forest composition across the landscape may take 100 years or more in the absence of major disturbances. Model results from the Tree Atlas and LINKAGES (Chapter 5) indicate substantial changes in habitat suitability or establishment probability for many species on the landscape, but do not account for migration constraints or differences among age classes. Forest landscape models such as LANDIS PRO can incorporate spatial configurations of current forest ecosystems, seed dispersal, and potential interactions between native species and the invasion and establishment of nonnative plant species (He et al. 1999, 2005). In addition, forest landscape models can account for differences among age classes, and have generally found mature trees to be more tolerant of warming (He and Mladenoff 1999). Because mature trees are expected to remain on the landscape, and recruitment of new species is expected to be limited, it is not expected that major 144 shifts in species composition will be observed by the middle of the century, except in areas that undergo more intensive harvesting or major stand-replacing disturbance events (Ryan et al. 2008). Climate change is projected to increase the intensity, scope, or frequency of some stand-replacing events such as wildfire, ice storms, and insect outbreaks, promoting major shifts in species composition where these events occur. Climate change is expected to affect early growth and regeneration conditions (medium evidence, medium agreement). Seedlings are more vulnerable than mature trees to changes in temperature, moisture, and other seedbed and early growth requirements. Evidence of climate change impacts on forest ecosystems is more likely to be seen in seedlings and early growth than in mature individuals. Temperature and moisture requirements for seed dormancy and germination are often much more critical than habitat requirements of an adult tree (Kitajima and Fenner 2000). Predicted changes in temperature, precipitation, growing season onset, and soil moisture may alter the duration or manifestation of germination conditions. For example, regeneration failure in balsam fir populations has been attributed, at least partially, to climate change (Abrams et al. 2001). For species with seeds that disperse successfully, these changes may result in redistribution on the landscape as seeds germinate only where conditions are met (Walck et al. 2011). Others species may fail to regenerate under altered future conditions, or may germinate without having sufficient conditions to develop. Warmer winters may promote the establishment of eastern redcedar and other southern species, although warmer temperatures alone are unlikely to drive their establishment (Abrams 2003). After establishment, advance regeneration (i.e., saplings) are still more sensitive than mature trees to drought, heat stress, frost, and other disturbances, such as fire, flooding, and herbivory (Kitajima and Fenner 2000). ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Increased fire frequency and harvesting will accelerate shifts in forest composition across the landscape (medium evidence, medium agreement). Studies from other regions show that increased fire frequency can accelerate the decline of species negatively affected by climate change and can accelerate the northward migration of southern tree species. Days with conditions that are suitable for wildfire ignition are expected to become more frequent, although the occurrence of wildfire will depend on both ignition and human response (Chapter 5). Frequent, low-intensity fires in certain forest ecosystems (e.g., beech-maple) can reduce or inhibit seedling establishment of tree species projected to decline under climate change (e.g., sugar maple, American beech). In other forest ecosystems (e.g., dry-mesic oak), fire can be beneficial for restoration and to promote regeneration. In some ecosystems (e.g., dry oak-pine), infrequent, highintensity fires can promote regeneration and release growing space for tree species that may be better adapted to future conditions. Fire (including lowintensity prescribed fire) is expected to accelerate changes in forest composition, promoting faster changes than those caused by increased temperature or moisture availability (He et al. 2002, Shang et al. 2004). Net change in forest productivity is expected to be minimal (medium evidence, low agreement). A few studies have examined the impact of climate change on forest productivity, but they disagree on how multiple factors may interact to influence it. Changes in productivity will likely be mixed and localized. Increases in drought, invasive plants, insects, disease, and wildfire are expected to negatively affect forest productivity in some parts of the region (Hanson and Weltzin 2000). Longer growing seasons, with adequate moisture, may lead to greater annual productivity. Future increases in carbon dioxide may enhance growth rates and water use efficiency of some species through carbon dioxide fertilization (Ainsworth and Rogers 2007, Norby et al. 2005), potentially offsetting the effects of drier growing seasons. Sulfur dioxide, a component of acid deposition, has been shown to reduce carbon dioxide fertilization effects in eastern redcedar in West Virginia (Thomas et al. 2013). Decreases in sulfur dioxide after the Clean Air Act of 1970 are allowing a slow recovery, which is expected to result in increased carbon uptake by trees. There is already some evidence for increased forest growth in the eastern United States (Cole et al. 2010, McMahon et al. 2010, Pan et al. 2009), but it remains unclear if long-term enhanced growth can be sustained (Bonan 2008, Foster et al. 2010). Nutrient and water availability, ozone pollution, and tree age and size all play major roles in the ability of trees to capitalize on carbon dioxide fertilization (Ainsworth and Long 2005). Productivity in the Central Appalachians is already affected by acid deposition, especially in those forests at the highest elevations (Elliott et al. 2013). Modeling results from LANDIS PRO, which do not include the possible effects of carbon dioxide fertilization or reductions in acid deposition, project minimal changes in basal area across the assessment area, even for GFDL A1FI, but large changes for some species in certain locations (Appendix 4). Elevation and aspect, and their influence on soil water availability, will also be a major driver of local ecosystem response (Vanderhorst et al. 2008). Adapive Capacity Factors Adaptive capacity is the ability of a species or ecosystem to accommodate or cope with potential climate change impacts with minimal disruption (Glick et al. 2011). Below, we summarize factors that could reduce or increase the adaptive capacity of forest ecosystems within the assessment area. Greater adaptive capacity tends to reduce climate change vulnerability, and lower adaptive capacity tends to increase vulnerability (Appendix 5). 145 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Low-diversity ecosystems are at greater risk (medium evidence, high agreement). Studies have consistently shown that diverse ecosystems are more resilient to disturbance, and low-diversity ecosystems are more vulnerable to change. In general, species-rich ecosystems have exhibited greater resilience to extreme environmental conditions and greater potential to recover from disturbance than less diverse ecosystems (Tilman 199, 1999). Consequently, less diverse ecosystems are inherently more susceptible to future changes and stressors (Swanston et al. 2011). Conversely, ecosystems that have low species diversity or low functional diversity (where multiple species occupy the same niche) may be less resilient to climate change or its associated stressors (Peterson et al. 1998, Walker 1992, Walker et al. 1999). Forest stands with low diversity of species, age classes, and genotypes have been more vulnerable to insect and disease outbreaks than diverse stands (Raffa et al. 2008). Genetic diversity within species is also critical for the ability of populations to adapt to climate change, because species with high genetic variation are more apt to have individuals that can withstand a wide range of environmental stressors (Reusch et al. 2005). Species in fragmented landscapes will have less opportunity to migrate long distances in response to climate change (limited evidence, high agreement). Evidence suggests that species may not be able to disperse over the distances required to keep up with climate change, but little research has been done in the region on this topic. Habitat fragmentation can hinder the ability of tree species to migrate to more suitable habitat on the landscape, especially if the surrounding area is nonforested (Ibáñez et al. 200, Iverson et al. 2004). It is estimated that a plant would need to migrate 90 miles north or 550 feet in altitude in order to escape a 1.8 °F increase in temperature (Jump and Peñuelas 2005). Modeling results indicate that 14 suitable habitat for tree species will migrate between 0 and 350 miles by the year 2100 under a high emissions scenario and between 30 and 250 miles under milder climate change scenarios (Iverson et al. 2004). Based on gathered data of seedling distributions, it has been estimated that tree species could possibly migrate northward at a rate of up to 100 miles per century (Woodall et al. 2009), and other evidence indicates that natural migration rates could be far slower for some species (Iverson et al. 2004a, McLachlan et al. 2005, Murphy et al. 2010). This research also suggests that range migration has already begun; centers of seedling densities were often more than 12 miles north of species’ centers of biomass (Woodall et al. 2009). Fragmentation makes migration even more challenging, because the landscape is essentially less permeable to migration (Jump and Peñuelas 2005, Scheller and Mladenoff 2008). Ecosystems that are highly limited by hydrologic regime or geological features may be topographically constrained (limited evidence, medium agreement). Our current understanding of the ecology of Central Appalachians ecosystems suggests that some species will be unable to migrate to new areas due to topographic constraints. Communities that require specific hydrologic regimes, unique soils or geology, or narrow elevation ranges may not be able to shift across the landscape, even if conditions are favorable. For example, high-elevation spruce/fir ecosystems are found exclusively in the highest elevations of the Allegheny Mountains, as remnant populations surviving in the coolest and wettest habitats in the region (Byers et al. 2007). These ecosystems, which range from wetlands to uplands, are already restricted to the highest elevations, and if habitat becomes unsuitable, it is doubtful that there will be alternate sites or that they would be able to migrate over unsuitable habitat to reach potential northern sites (Nowacki et al. 2009). ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Ecosystems that are tolerant of disturbance or are disturbance-adapted have less risk of declining on the landscape (medium evidence, high agreement). Basic ecological theory and other evidence support the idea that systems that are adapted to more frequent disturbance will be at lower risk. Disturbances such as drought, flooding, wildfire, and insect outbreaks have the potential to increase in the assessment area (Chapters 4 and 5). Several ecosystems (e.g., Appalachian [hemlock]-northern hardwood and north-central interior beechmaple forests) are adapted to frequent gap-phase disturbances, but undergo stand-replacing events on the scale of hundreds or thousands of years. Therefore, these systems may be less tolerant of more frequent stand-level disturbances, such as drought or fire. Mesic ecosystems can create conditions that could buffer against fire and drought to some extent (Nowacki and Abrams 2008). However, even species in mesic ecosystems could decline if soil moisture declines significantly. Forest ecosystems that are more tolerant of drought, flooding, or fire (e.g., dry oak and oak/pine forest and woodland) will likely be better able to withstand climate-driven disturbances (Wagner et al. 2012). This principle holds true only up to a point, because it is also possible for disturbance-adapted ecosystems to undergo too much disruption. For example, oak and pine ecosystems might cover a greater extent under drier conditions with more frequent fire, but these systems might also convert to barrens or open grasslands if fire becomes too frequent or drought becomes too severe. Fire-adapted ecosystems will be more resilient to climate change (high evidence, medium agreement). Studies have shown that fire-adapted ecosystems are better able to recover after disturbances and can promote many of the species that are expected to do well under a changing climate. In general, fire-adapted ecosystems that have a more open structure and composition are less prone to high-severity wildfire (Shang et al. 2004). Frequent low-severity fire has also been shown to promote many species projected to do well under future climate projections, such as shortleaf pine, pitch pine, and a number of oak species (Aldrich et al. 2010, Brose and Waldrop 200, Brose et al. 2013, Patterson 200). In these ecosystems, fire suppression has resulted in sometimes heavy encroachment of woody species in the understory that compete with oak and pine regeneration (Nowacki and Abrams 2008, Patterson 200, Sharitz et al. 1992). In addition, fire suppression in fire-adapted ecosystems can lead to increased susceptibility to damaging insect infestations (McCullough et al. 1998). Since the mid-1900s, suppression of fire has led to an increase in red maple and sugar maple across the eastern forests (Abrams 1998, Nowacki and Abrams 2008). These species are not projected to fare well under climate change, largely due to regeneration failure (Chapter 5), and the opportunity may arise to restore firesuppressed ecosystems. However, the effects of fire on seedling establishment, tree growth, and nutrient cycling can vary by site conditions, species, and burn regime (Brose et al. 2013, McCullough et al. 1998). Ecosystems occupying habitat in areas of high landscape complexity have more opportunities for persistence in pockets of refugia (medium evidence, medium agreement). The diversity of landscape positions occupied by forest may provide opportunities for natural refugia, for example where cool air and moisture accumulate in valley bottoms. Species diversity in the Central Appalachians has been linked to geophysical diversity of the area (Anderson and Ferree 2010). With increasing topographic and landform complexity come a greater number of landscape characteristics and microhabitats that buffer against climate changes 147 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES (Anderson et al. 2012, Fridley 2009). Many areas across West Virginia and Maryland, such as the Allegheny Mountains, have a high diversity of landscape characteristics, such as geophysical setting, landscape complexity, and connectivity, that contribute to the high species diversity (Anderson et al. 2012). This diversity of landscape features supports a variety of rare, endemic, and localized plant and animal species, some of which are restricted to a single geology (Anderson and Ferree 2010). The Tree Atlas modeled the most common tree species, but did not model many of the rare species (Chapter 5). Even the relatively flat areas of the assessment area contain complex ridge systems and associated soil moisture regimes that support a high diversity of species. Although climate will largely determine a species’ potential range, it is the influence of geology that creates areas of microhabitat offering refugia against the effects of climate change (Anderson and Ferree 2010). VuLNERABiLiTY DETERMiNATioNS FoR iNDiViDuAL FoREST ECoSYSTEMS Climate-induced shifts in drivers, stressors, and dominant tree species will result in different impacts to forest ecosystems within the assessment area. Some ecosystems may have a greater capacity to adapt to these changes than others, whereas some may be susceptible to relatively minor impacts. Therefore, it is helpful to consider these factors for individual forest ecosystems in addition to describing general principles related to vulnerability and adaptive capacity. Table 23 presents a summary of current major drivers and stressors for each forest ecosystem covered in this assessment. The following vulnerability determinations draw on the information presented in previous chapters, as Table 23.—Vulnerability determinaion summaries for forest ecosystems considered in this assessment Potenial impacts Adapive capacity Vulnerability Evidence Agreement Negaive Low-Moderate High Medium Medium Dry calcareous forest, woodland, and glade Neutral-Negaive Low-Moderate Moderate-High Limited-Medium Medium Dry oak and oak/pine forest and woodland Posiive Moderate-High Low Medium Medium-High Dry/mesic oak forest Posiive-Neutral High Low- Moderate Medium Medium-High Large stream loodplain and riparian forest Negaive Low High Medium Medium Mixed mesophyic and cove forest Neutral-Negaive Moderate-High Moderate Limited-Medium Medium North-central interior beech/maple forest Neutral Moderate Moderate Limited-Medium Medium Small stream riparian forest Negaive Moderate Moderate-High Medium Medium Spruce/ir forest Negaive Moderate High Limited-Medium Medium Forest ecosystem Appalachian (hemlock)/ northern hardwood forest 148 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES well as an expert panel assembled from a variety of organizations and disciplines across the assessment area. The 19 panelists evaluated anticipated climate trends for the assessment area and ecosystem model projections (Chapter 5), in combination with their own expertise. For each forest ecosystem, panelists considered the potential impacts and adaptive capacity to assign a vulnerability determination (Fig. 43) and a level of confidence in that determination using the same confidence scale described above. For a complete description of the methods used to determine vulnerability, see Appendix 5. Overall vulnerability determinations were rated lowest for dry oak and oak/pine forest and woodland and highest for Appalachian (hemlock)/northern hardwood, spruce/fir, and large stream floodplain and riparian forests (Table 23). Impacts were rated as being most negative for Appalachian (hemlock)/ northern hardwood, large stream floodplain and riparian, and spruce/fir forests. Impacts were rated most positive for dry oak and oak/pine forest. Several negative and positive impacts were identified for north-central interior beech/maple forest, which was given an overall rating of neutral impacts. Adaptive capacity was rated lowest for large stream floodplain and riparian forest, and highest for dry/mesic oak forest. Panelists tended to rate the amount of evidence as limited to medium (between limited and robust) for most forest ecosystems. Incomplete knowledge of future wildfire regimes, interactions among stressors, and precipitation regimes were common factors limiting this component of overall confidence. The ratings of agreement among evidence also tended to be in the medium range. Contrasting information related to precipitation regimes under Figure 43.—Vulnerability diagram used in the assessment. the high and low climate change scenarios was one factor that limited the level of agreement among evidence. The way that forest ecosystems were organized and described for this assessment also limited the agreement in some instances. In general, ratings were slightly higher for agreement than for evidence. Evidence appears not to be as robust as the experts would prefer, but the various information sources that are available tend to support similar conclusions. In the sections that follow, we summarize the climate-related impacts on drivers, stressors, and dominant tree species that were major contributors to the vulnerability determination for each forest ecosystem. In addition, we summarize the main factors contributing to the adaptive capacity of each ecosystem. For a list of common tree species in each forest ecosystem, see Chapter 1. 149 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Appalachian (Hemlock)/Northern Hardwood Forest High Vulnerability (medium evidence, medium agreement) Climate change may intensify several interacing stressors, such as drought, forest pests, and invasive species. Any increased wildire acivity would be detrimental to the health of this forest type. Reduced species diversity may decrease resilience to the future climate. Valley botoms and other microsites in areas of complex topography may be bufered from the efects of climate change, providing refugia. Negaive Potenial Impacts Drivers: Decreased precipitation and increased temperatures may interact to ultimately decrease soil moisture during summer and fall. Increased heat and moisture stress, exposure to insect pests and pathogens, and more frequent disturbances are expected to interact and place increased stress on this ecosystem. Climate change may also alter the gap-phase dynamics that enable the regeneration of many shade-tolerant species if damaging storms, pest outbreaks, or wildfires become more frequent or widespread. Dominant Species: Model results indicate that American beech, eastern hemlock (considered a keystone species where it occurs), and sugar maple will remain relatively stable for PCM B1, but will lose suitable habitat, growth potential, and volume in the assessment area for GFDL A1FI (Chapter 5). These species are vulnerable to the direct changes in temperature and precipitation, and are susceptible to moisture stress, beech bark disease, hemlock woolly adelgid, and other climate and nonclimate stresses. Results are mixed for red maple, tulip tree, black cherry, and white ash, which are projected to lose suitable habitat but maintain potential growth and volume. Although the amount of suitable habitat may contract, models agree that remaining suitable habitat may allow regeneration of these species in the absence of other stressors. Red spruce, considered a minor component in the eastern part 150 of the assessment area, is projected to experience a dramatic decline in growth potential and suitable habitat, especially for GFDL A1FI, although established adults are likely to persist even after they no longer regenerate successfully. Stressors: Climate change may amplify several major stressors to northern hardwoods, particularly for stands occurring on southwest slopes or marginally suitable soils. Hemlock woolly adelgid, beech bark disease, emerald ash borer, and other pests currently attack many species in this ecosystem and may cause more frequent and severe damage in climate-stressed forests. Pests such as Asian longhorned beetle and gypsy moth may present new risks if they are able to expand from established locations adjacent to the assessment area (Chapter 5). With the exception of spruce/fir forests, acid deposition afflicts this ecosystem more than any other due to its distribution across acid-sensitive geologies (USDA 200). Acid deposition damages ecosystem health, and it is unclear how climate change may affect the ability of ecosystems to cope with acid deposition in the future. Increases in wildfire risk would be a severe impact for this ecosystem because many tree species within this ecosystem do not tolerate fire. Interactions between stressors, such as drought, pests, acid deposition, invasive species, and wildfire are likely to have greater impacts than temperature or precipitation alone. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Low-Moderate Adapive Capacity This ecosystem typically supports a variety of plant species on gentle to steep slopes in diverse and complex terrain, but is limited to the highest elevations in the Allegheny Mountains. This ecosystem is often found on soils with high waterholding capacity in areas that normally receive abundant moisture from precipitation and ground seepage. However, the combined effects of acid deposition, drought, and defoliation have already resulted in lower species diversity and reduced adaptive capacity (Chapter 5). Eastern hemlock is currently susceptible to widespread mortality A hemlock stand in Wooster Memorial Park, Ohio. Photo by David M. Hix, Ohio State University, used with permission. from hemlock woolly adelgid, which is expected to dramatically reduce eastern hemlock populations over the next few decades. Red spruce is currently expanding on the landscape, and may persist where cool, wet conditions provide refugia. Sites on moist soils that continue to receive abundant moisture may be buffered from seasonal moisture stress, whereas sites on exposed slopes may be more sensitive to moisture stress. The diversity of landscape positions occupied by this forest may also provide opportunities for natural refugia, for example, where cool air and moisture pool in north-facing pockets and valley bottoms. An Appalachian (hemlock)/northern hardwood forest. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. An Appalachian (hemlock)/northern hardwood forest. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 151 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Dry Calcareous Forest, Woodland, and Glade Moderate-high Vulnerability (limited-medium evidence, medium agreement) Many of the species in this ecosystem are projected to do well under a range of plausible climate futures. However, this ecosystem’s occupaion of extreme habitat on unique soil types implies that it is geographically limited, and unable to shit freely on the landscape. Neutral-Negaive Potenial Impacts Drivers: Drought and ire have been important factors in maintaining the open form of this ecosystem, and increased drought is not expected to harm many species, unless it becomes too severe. Wildire potential could increase under drier conditions, although ire intensity will determine whether it is a positive or negative impact. Lowintensity ire could beneit this ecosystem by reducing the eastern redcedar component in the woodland portions of this ecosystem that are becoming overgrown. However, high-intensity ire that results in widespread mortality of eastern redcedar will dramatically change this ecosystem. Dominant Species: Projected changes in climate are expected to beneit many of the common tree species in this ecosystem. Models project that eastern redcedar, white oak, and post oak will remain relatively stable or increase in suitable habitat, potential growth, and volume under both climate scenarios (Chapter 5). White oak is longlived and able to persist in the shaded understory until openings are naturally created in the canopy. Eastern redcedar, and to a lesser extent white oak, is dependent on disturbance and expected to beneit from soil moisture deicits, ire, and other disturbances. Chinkapin oak, eastern redbud, eastern hophornbeam, and shagbark hickory were modeled only by the Tree Atlas, and were similarly projected to increase in suitable habitat. 152 Stressors: Increased drought duration and extent may increase susceptibility to oak decline, or may combine with insect and disease factors to increase stress or mortality. Invasive species are also common in this forest ecosystem, and climate change is expected to promote establishment and growth of invasives, resulting in increased competition with the many rare native plants in this ecosystem. Increases in invasive species such as cheat grass, stilt grass, and bush honeysuckles could increase ire fuels in this type, leading to potentially more-intense ire when it does occur. Low-Moderate Adapive Capacity This ecosystem is characterized by high species diversity, but has been affected by limestone quarrying, agriculture, and fragmentation. The cooccurrence of the dominant species is tightly linked to the unique soils derived from limestone, and movement on the landscape is limited to landscape positions where those soils form. Many of these species tolerate temperature and moisture extremes, especially on exposed landscape features, allowing them to outcompete other species. Severe drought, projected temperature increases, and increased ire may allow expansion of the woodland and glade elements at the expense of the forested areas. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES A north-central Appalachian circumneutral clif and talus ecosystem. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry calcareous outcrop at Castle Rock, West Virginia. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry calcareous woodland. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 153 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Dry Oak and Oak/Pine Forest and Woodland Low Vulnerability (medium evidence, medium-high agreement) This ecosystem is the most resilient to heat and drought, with many of the species currently doing well, and projected to do well under future climate. Periodic condiions that limit regeneraion may be bufered by oak’s ability to resprout. Increased drought and ire are likely to beneit this ecosystem, discourage invasive species, and maintain an open structure that promotes oak and pine regeneraion. Posiive Potenial Impacts Drivers: This dry ecosystem is characterized by thin, droughty, and nutrient-poor soils. Soils can become hydrophobic for short periods of time, which can be made worse by fire. Fire frequency was historically higher than it is currently, largely due to fire suppression over the last 50 years. Drier soil conditions in summer and fall, especially on south-facing slopes, may increase the risk of wildfire (Chapter 5). Shale barrens and ridge tops are especially exposed to extreme temperatures. Increased frequency of extreme weather events (e.g., windstorms and ice storms) may lead to more frequent disturbances. Dominant Species: Many of the common species in this ecosystem are projected to remain relatively stable in total volume, but volume is expected to shift from many smaller trees in younger age classes to fewer larger trees in older age classes. Many of the species in this ecosystem will require active management, such as prescribed fire, to stimulate regeneration. Models project that suitable habitat, potential growth, and trees per acre for chestnut oak and scarlet oak will remain stable for PCM B1 and decrease for GFDL A1FI. Black oak is projected to remain stable for PCM B1, but for GFDL A1FI 154 suitable habitat is expected to increase while growth potential and trees per acre decrease. Models project increases in suitable habitat and potential growth for only one species, loblolly pine, which is expected to benefit from increased temperatures under both scenarios. Pitch pine, Table Mountain pine, and Virginia pine were modeled only by the Tree Atlas, which projected suitable habitat to remain stable or increase for both scenarios. Stressors: Increased drought conditions, especially during the growing season, may increase susceptibility to red oak borer, gypsy moth, armillaria root disease, and other insect pests and diseases. Southern pine beetle outbreaks have been observed in New Jersey and Pennsylvania systems recently, and may increase due to warmer temperatures. Ailanthus, Japanese stiltgrass, multiflora rose, bush honeysuckle, autumn olive, and Japanese barberry often outcompete native herbs and shrubs in this ecosystem, and are also likely to benefit from warmer temperatures and increased disturbance. Invasive shrubs in the understory may provide additional ladder fuels and increase fire intensity where wildfire occurs; impacts will depend on severity of fire. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Moderate-High Adapive Capacity This ecosystem is the most resilient to extreme heat and moisture deficits of the ecosystems examined in the assessment area. Many pine and oak species are fire adapted and drought-tolerant, some requiring high-intensity fire to regenerate (Vose et al. 1993). A history of fire suppression and succession has contributed to a reduced pine component in favor of oak species. Increased wildfire frequency could help regenerate and promote both oak and pine species. Low-severity late-season drought generally favors oak species, although severe drought may hinder regeneration, or combine with other stressors to make individuals more susceptible to mortality or reduced productivity. This ecosystem benefits from disturbance regimes, such as fire and windthrow, which promote conditions for regeneration. This ecosystem’s wide distribution on a range of habitat conditions makes it well-poised to take advantage of new habitat that may become too dry for other species. A dry oak forest ecosystem with rhododendron in the understory. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry oak forest ecosystem with an open canopy. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry oak forest ecosystem. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. An oak/pine woodland. The understory of this dry site is dominated by sedge and grasses. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 155 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Dry/Mesic Oak Forest Low-Moderate Vulnerability (medium evidence, medium-high agreement) This ecosystem supports a high number of tree species and occurs over a wide range of habitats. Many species are tolerant of dry soil condiions and ire, although young regeneraion may be sensiive to severe drought and ire. Southern oak and hickory species are likely to beneit from projected changes in climate. Posiive-Neutral Potenial Impacts Drivers: Fire frequency was historically higher than it is currently, largely due to fire suppression over the last 50 years. Drier soil conditions in summer and fall, especially on south-facing slopes, may increase the risk of wildfire. Increased frequency of extreme weather events (e.g., windstorms and ice storms) may lead to more frequent large-gap disturbances. Increases in extreme precipitation events may increase the potential for erosion and channeling. Dominant Species: Of the many species modeled, suitable habitat was generally projected to increase for the southern oaks and hickories, whereas other common species are projected to persist over a smaller extent. Models project that habitat suitability, basal area and trees per acre, and potential growth for pignut hickory and white oak will remain relatively stable or increase slightly under both scenarios. Results for northern red oak are highly variable across the assessment area, but suggest positive effects on regeneration where suitable habitat remains. Other common species are not expected to do as well, especially for GFDL A1FI: models project that suitable habitat, potential growth, and trees per acre will decrease for chestnut oak and scarlet oak. Black oak is projected to remain stable for PCM B1, but for GFDL A1FI suitable habitat is expected to increase while growth potential and trees per acre decrease. Mockernut hickory and shagbark hickory were modeled only by the Tree Atlas, and both are projected to increase in suitable habitat. 15 Stressors: Increased drought risk, especially during the growing season, may increase susceptibility to red oak borer, ambrosia beetle, gypsy moth, armillaria root disease, and other insect pests and diseases. Ailanthus, Japanese stiltgrass, and garlic mustard, which often outcompete native herbs and shrubs in this ecosystem, are expected to do well in warmer temperatures. Low-severity late-season drought generally favors oak species, although severe drought may hinder regeneration, or combine with other stressors to make individuals more susceptible to mortality or reduced productivity. High Adapive Capacity A history of fire suppression and timber harvesting has facilitated a shift to more mesic soils and associated hardwood species (e.g., sugar maple, American beech, tulip tree). Increased fire frequency could help regenerate oak species and restore the understory composition. However, very frequent fires have the potential to kill young seedlings of any species, even those species that have relatively fire-resistant, thick bark as adults. This ecosystem is widely distributed, representative of a range of habitat conditions, and likely to expand on the landscape. American chestnut was historically a dominant canopy tree but now cannot grow past sapling size due to chestnut blight. Blight-resistant American chestnut variants are currently under development and experimental planting is already occurring, resulting in increased species diversity in select areas (Jacobs et al. 2013). ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES A mesic oak forest with maple regeneraing in the understory. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry oak forest with grasses dominaing the open understory. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A dry/mesic oak forest. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 157 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Large Stream Floodplain and Riparian Forest High Vulnerability (medium evidence, medium agreement) Climate change is expected to alter the water regimes in riparian systems, and may amplify the efects of insect pests, invasive species, and polluion. Dependence on periodic inundaion, combined with compeiion from invasive species, may result in a reduced ability of naive tree species to tolerate increased disturbances. Negaive Potenial Impacts Drivers: Potential changes to the precipitation regime could intensify peak streamflow and shift the timing to earlier in the spring. Reduced precipitation in the summer and fall would result in drier conditions, increasing the potential for latesummer drought. An increase in intense precipitation events is likely to result in more frequent flooding. Wildfire, currently episodic and human-caused, could increase under drier conditions, although the extent would be limited by the fragmented nature of riparian and floodplain ecosystems. Dominant Species: Many riverine species in this forest type were modeled only by the Tree Atlas; thus evidence is somewhat limited regarding dominant species. Black willow, green ash, sweetgum, and sycamore are projected to increase in suitable habitat over much of the assessment area. Silver maple had mixed results, but is projected to generally decrease in suitable habitat for PCM B1 and increase for GFDL A1FI. Eastern cottonwood and bur oak occurred at sufficient densities to be modeled only in the Ohio portion of the assessment 158 area, and are projected to decrease slightly for PCM B1 and increase for GFDL A1FI. Pin oak, also adequately abundant only in Ohio, is projected to increase and expand into West Virginia, where pin oak swamps currently exist in isolated locations. These species are all tightly linked to moisture availability, and are especially threatened by potentially drier soil conditions. Stressors: Climate change is expected to intensify several key stressors for large stream riparian and floodplain forests. Many invasive plant species currently threaten this ecosystem and are expected to benefit from climate change and outcompete native species. Drought-stressed trees may become more susceptible to insect pests such as emerald ash borer and diseases such as thousand cankers and elm yellows. Interactions among multiple stressors may also lead to more severe climate change impacts. Increases in storm intensity and flooding events have the potential to increase soil erosion and sedimentation, and compound anthropogenic stressors such as agricultural runoff and industrial pollution. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Low Adapive Capacity This ecosystem exists in many variations within a relatively small proportion of the assessment area, but is extremely altered by habitat destruction, fragmentation, and disconnection of floodplain forests from rivers and streams (e.g., by roads or other infrastructure that impedes the flow of water). The high number of invasive species outcompeting natives has already reduced the adaptive capacity of this ecosystem. Although this ecosystem is highly dependent on disturbance and a regular influx of A large stream loodplain forest on the Buckhanon River, West Virginia. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. seeds, nutrients, and water during periodic flooding, increases in flood intensity or more frequent drought may not be tolerated by many species, especially in the early growth stages. Mortality of ash species from emerald ash borer is likely to eliminate this species by mid-century, reducing overall native species diversity. Forests located along river corridors may be buffered from water deficit better than those located farther away on the flood plain, but will be more exposed to flooding effects. A large stream loodplain forest on the Meadow River, West Virginia. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A large stream loodplain forest on the Greenbrier River, West Virginia. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 159 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Mixed Mesophyic and Cove Forest Moderate Vulnerability (limited-medium evidence, medium agreement) This relaively sheltered ecosystem may face a suite of ampliied disturbances, including wildire, drought, and invasion by invasive species. Suitable habitat for many species is projected to decline, although there is great potenial for the complex topography to provide refugia where disjunct populaions may persist. Neutral-Negaive Potenial Impacts Drivers: This ecosystem is adapted to generally wet or mesic sites in the Allegheny Plateau and Allegheny Mountains sections, and is characterized by a high number of tree species. If drought becomes more frequent or widespread in late summer or fall, seedlings and saplings may be at risk of desiccation. Drought would lead to increased risk of wildfire, which this ecosystem would not tolerate well. Increased frequency of extreme weather events (e.g., windstorms and ice storms) may lead to more frequent large-gap disturbances. cherry, and white ash, which are projected to lose suitable habitat but maintain potential growth and volume. Although the amount of suitable habitat may contract, models agree that remaining suitable habitat may allow regeneration of these species in the absence of other stressors. Results for northern red oak are highly variable across the assessment area, but suggest positive effects on regeneration where suitable habitat remains. Black oak is projected to remain stable for PCM B1, but for GFDL A1FI suitable habitat is expected to increase while growth potential and trees per acre decrease. Dominant Species: Many species are commonly associated with this ecosystem, and individual species responses are expected to differ with ecological sections and expected degree of climaterelated changes. Models project that American beech, eastern hemlock (considered a keystone species where it occurs), and sugar maple will remain relatively stable for PCM B1, but will lose suitable habitat, growth potential, and volume in the assessment area for GFDL A1FI (Chapter 5). These species are vulnerable to the direct changes in temperature and precipitation, and are susceptible to moisture stress, beech bark disease, mortality from hemlock woolly adelgid, and other stresses resulting from indirect impacts of climate change. Results are mixed for red maple, tulip tree, black Stressors: Increased drought conditions may increase susceptibility of trees in this system to hemlock woolly adelgid, forest tent caterpillar, beech bark disease, and other insect pests and diseases. Eastern hemlock is currently susceptible to widespread mortality from hemlock woolly adelgid, which is expected to dramatically reduce eastern hemlock populations over the next few decades. Japanese stiltgrass, garlic mustard, ailanthus, and bush honeysuckle have already shifted understory species composition, and are expected to increase in response to warmer temperatures. Increases in invasive species could increase fire fuels in this type, leading to potentially more-intense fire when it does occur. Most species are fire-intolerant, although oak species would benefit from an increase in fire. 10 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Moderate-High Adapive Capacity This ecosystem currently has high species diversity, and its sheltered position on concave slopes in complex topography may buffer against climate changes. The ability of coves to collect water and nutrients from higher areas may benefit species by creating refugia from temperature increases, precipitation changes, and wind. Ecosystem response to climate change impacts will vary across the landscape depending on current landscape position, individual species response, and connectivity. In the mountains, species may be able to migrate upwards more easily than northwards to escape warming temperatures. Emerald ash borer infestations have already damaged and killed many ash trees. This forest ecosystem has been diminished by fragmentation and conversion to agriculture, coal mining, and logging. Especially in southeastern Ohio, remaining forest blocks occur in a highly fragmented mosaic of second-growth forests and have reduced biodiversity. A cove forest in the Allegheny Mountains of West Virginia. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A Southern and Central Appalachian cove forest. Photo by Brian Streets, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 11 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES North-Central Interior Beech/Maple Forest Moderate Vulnerability (limited-medium evidence, medium agreement) Although sugar maple and American beech are projected to decline to some degree, many associated species in this ecosystem are projected to do well under a range of future climates. This forest’s posiion on glacial ill loodplains, moraines, and plateaus promotes and preserves moist soil condiions, a criical feature which may help bufer the impacts of changing temperature and hydrologic regimes. Neutral Potenial Impacts Drivers: This forest occurs largely within the lakeeffect zone of Lake Erie, where heavy-textured soils and glacial landforms help retain soil moisture. Other instances are found on lowland positions supplied by wetland hydrology. Projected decreases in precipitation in late summer and fall may increase the frequency or extent of drought. This system is intolerant of fire, and is characterized by long disturbance intervals. Increased frequency of extreme weather events is likely to promote canopy gap disturbances of larger size and extent than at present. Dominant Species: Models project that American beech, sugar maple, and eastern hemlock (occurring locally in the glaciated Ohio and eastern portions of the assessment area) will remain relatively stable for PCM B1, but will lose suitable habitat, growth potential, and volume in the assessment area for GFDL A1FI (Chapter 5). These species are vulnerable to the direct changes in temperature and precipitation, and are susceptible to increased moisture stress and other indirect impacts of climate change. Results are mixed for red maple, tulip tree, black cherry, and white ash, which are projected to lose suitable habitat but maintain potential growth and volume. Although the amount of suitable habitat may contract, models agree that remaining suitable habitat may allow regeneration of these species in the absence of other stressors. Results for northern red oak are highly variable across the assessment area, but suggest positive effects on regeneration where suitable habitat remains. 12 Stressors: Beech bark disease, emerald ash borer, hemlock woolly adelgid, anthracnose disease, and a variety of other pests and pathogens currently affect this ecosystem. Certain insects, such as hemlock woolly adelgid, may benefit from warmer winter temperatures, creating additional stress for these forests. The emerald ash borer has already reduced the white ash component in parts of the assessment area. Invasive plants such as princesstree, silktree, ailanthus, and glossy buckthorn compete directly with understory plants and native tree regeneration and these invasives are likely to take advantage of increased temperatures and disturbance. Moderate Adapive Capacity This ecosystem supports relatively high species diversity. Its position on moist soils in glacial topography, and its proximity to lake-effect precipitation, helps to maintain soil moisture, which may buffer against drought and discourage conditions that promote wildfire. However, these benefits decrease with increasing distance from Lake Erie. Many of the dominant tree species are not tolerant of drought or fire. Drought-stressed trees may be more susceptible to invasives or disease complexes, resulting in decreased productivity or mortality. An increase in wildfire could promote transition to primarily fire-adapted species (e.g., oaks), changing the identity of this ecosystem. Heavy deer browsing is also limiting seedling establishment and growth, and protection from herbivory will be critical in establishing regeneration, now and under future climate conditions. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Sugar maple and beech canopy in an Ohio beech/maple forest. Photo by David M. Hix, Ohio State University, used with permission. A north-central interior beech/maple forest at Crall Woods, Ohio. Photo by David M. Hix, Ohio State University, used with permission. 13 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Small Stream Riparian Forest Moderate-High Vulnerability (medium evidence, medium agreement) This ecosystem is adapted to natural disturbance, but is threatened by ampliicaion of the disturbance regime, and by invasive plants, insects, and pathogens. Many species are projected to remain stable or increase under a range of future climate condiions, but a keystone species, hemlock, is likely to disappear in many areas. Negaive Potenial Impacts Drivers: Changes to the timing and intensity of precipitation events may lead to increased flashiness and more frequent high water events in spring. Spring flooding and inundation have the potential for increased erosion, silt loads, and sedimentation. Summer and fall moisture deficits have the potential to create dry vegetation conditions, stressing hydrophilic seedlings and supporting wildfire conditions. Mortality and damage from drought or storms may result in increased coarse woody debris, contributing to wildfire fuels. Dominant Species: Many riverine species in this forest type were modeled only by the Tree Atlas, so evidence is somewhat limited regarding dominant species (Chapter 5). Additionally, some of these species are not common on the landscape, and are therefore difficult to model. Suitable habitat is projected to remain stable or increase for sycamore, river birch, black walnut, and boxelder. Silver maple and cottonwood are projected to decrease for PCM B1 and increase for GFDL A1FI. Hemlock and red maple were modeled by all three models. Eastern hemlock is projected to remain stable or decrease in suitable habitat and potential growth; basal area and trees per acre are projected to decrease due to succession, and to a lesser extent due to climate. Red maple had mixed results for suitable habitat and potential growth, and basal area and trees per acre are projected to increase due to succession and climate change. Many of these species, except red maple, are tightly linked to moisture availability. 14 Stressors: Invasive plants are very problematic in this ecosystem, with greater impacts generally occurring downstream. Increased flashiness followed by dry periods could cause amplification of the current hydrologic cycle, potentially increasing the spread and establishment of current and newly introduced invasive species. Drought-stressed trees may be more susceptible to diseases such as thousand cankers and elm yellows, and insect pests such as hemlock woolly adelgid. Increases in storm frequency and flood intensity have the potential to increase soil erosion and sedimentation, and compound anthropogenic stressors such as agricultural runoff and industrial pollution. Moderate Adapive Capacity This ecosystem exists in many variations or settings across the landscape with various assemblages of a fairly diverse set of species, many of which are projected to remain stable or even increase under climate change. Further, this ecosystem type is adapted to cope with a high level of variability and natural disturbance, and may be able to handle many impacts of temperature and precipitation changes except for extreme drought or severe flooding. Cold air pooling in valleys and shelters may also provide refugia that are buffered from temperature increases. In the cooler and moister sites, hemlock is a keystone species that has been declining and is projected to decline further. For these forests, the loss of hemlock is likely to change the species assemblage dramatically, with fast-growing generalists like red maple or a variety of invasive species likely to overtake the newly vacated niche. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES A small stream riparian forest with a large herbaceous component. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. A small stream riparian forest. Photo by Jim Vanderhorst, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. Cotonwood and other hardwoods along a stream at Alum Creek, Ohio. Photo by David M. Hix, Ohio State University, used with permission. 15 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Spruce/Fir Forest High Vulnerability (limited-medium evidence, medium agreement) This ecosystem is dependent on very moist condiions, and persists only in the coolest, wetest, and highest elevaion sites in mountainous secions. Projected increases in temperature and decreases in summer and fall precipitaion may exceed the ecological tolerances of this ecosystem’s deining species. Complex topography may provide cool pockets of habitat where these species would be likely to persist. Negaive Potenial Impacts Drivers: This ecosystem type is adapted to cool temperatures and abundant moisture in the form of rain, snow, and fog drip. Projected increases in temperature and potential decreases in precipitation later in the growing season may decrease the amount of atmospheric moisture that could develop into fog in summer or snow in winter. If soils in this ecosystem dry out, the entire ecosystem would be affected. Drier conditions could also increase the risk of duff fire, previously not a threat except after extreme anthropogenic disturbances. Changes in winter processes could affect this high-elevation ecosystem more than others; interacting effects of reduced snow cover (warm temperatures) or increased snow cover (lake effect) may alter soil freezing conditions. Dominant Species: Red spruce and balsam fir (the two keystone species in this ecosystem) are limited to the Allegheny Mountains and the Northern Ridge and Valley sections, and models project suitable habitat and growth potential to decline dramatically for both species under both climate scenarios (Chapter 5). Models also project suitable habitat, growth potential, and trees per acre to decline for 1 eastern hemlock and eastern white pine, but only for GFDL A1FI. Results are mixed for red maple, tulip tree, black cherry, and white ash, which are projected to lose suitable habitat but maintain potential growth and volume. Although the amount of suitable habitat may contract, models agree that remaining suitable habitat may allow regeneration of these species in the absence of other stressors. Other common species were modeled only by the Tree Atlas: cucumbertree, yellow birch, and sweet birch are also projected to lose suitable habitat in the sections occupied by this ecosystem. Stressors: Insect pests such as the hemlock and balsam woolly adelgids currently affect this ecosystem and have the potential to increase when winter temperatures no longer limit populations. There is also potential for new invasive plants, although they may be limited by acidic soils. If deer populations benefit from warmer temperatures, herbivory on hemlock and balsam fir could increase, but red spruce would benefit because it is not a preferred browse species. Acid deposition damages ecosystem health, and it is unclear how climate change may affect the ability of ecosystems to cope with acid deposition in the future. ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES Moderate Adapive Capacity This forest ecosystem is currently stable and expanding on the landscape to reoccupy available suitable habitat. Red spruce, projected to decline for GFDL A1FI at the end of the century, influences the soil to create positive edaphic conditions that are favorable to its own regeneration. Red spruce has been negatively affected by acid deposition, which may decrease its natural resistance to changes (McLaughlin and Kohut 1992, McLaughlin et al. 1990, Schuler and Collins 2002). Balsam fir has the lowest adaptive capacity of all the species in this ecosystem, largely due to its fire- and droughtintolerance and susceptibility to balsam woolly adelgid and other insect pests. Eastern hemlock is currently susceptible to widespread mortality from hemlock woolly adelgid, which is expected to dramatically reduce eastern hemlock populations over the next few decades. The potential for drought may be buffered by high rainfall and fog generated at higher elevations. Suitable habitat for this ecosystem is already limited to the highest elevations in the Central Appalachians and the range of this ecosystem may contract as climate change forces species upward. Cold air pooling in valleys and shelters may provide areas of refugia buffered from temperature increases. Red spruce is currently expanding on the landscape, and may persist where cool, wet conditions provide refugia. A high-elevaion spruce/ir forest in West Virginia. Photo by Elizabeth Byers, West Virginia Division of Natural Resources, Natural Heritage Program, used with permission. 17 ChAPTER 6: FoREST ECoSYSTEM VuLNERABiLiTiES ChAPTER SuMMARY Forest ecosystems across the assessment area will be affected by climate change, although these ecosystems and individual tree species will respond to these changes differently. The synthesis statements in the first half of this chapter can be applied as general principles when specific information about expected climate change impacts is lacking. Overall, we expect that forest ecosystems will be most severely affected by projected decreases in late season precipitation; decreases are projected for summer for GFDL A1FI and for fall for PCM B1. Forest ecosystems that are adapted to dry conditions and frequent disturbances are expected to be less vulnerable to the range of future climates. Forest ecosystems that are adapted to tolerate a wide range of conditions and disturbances, and have higher mobility on the landscape, are also expected to be better able to persist under a range of plausible climates. The vulnerability determinations for individual forest ecosystems are best interpreted as broad trends and expectations across the assessment area. For some species, climate-related changes over the next century may be a continuation of current trends. For other species, it may take more than 100 years before such changes become apparent. For long-lived species especially, substantial changes on the landscape within this century will likely be the result of succession, management, and disturbance. Vulnerability to anthropogenic stressors such as fragmentation, urban development, and arson impinges on an ecosystem’s adaptive capacity, and may be much more influential on ecosystems than climate change, especially over the first half of this century. This assessment uses the most up-to-date information from the scientific literature, a coordinated set of ecosystem modeling results and climate projections, and the input of a large team 18 of local experts. Even so, there are limitations and unknowns that make these determinations imperfect. As new information continues to be generated on the potential impacts of climate change on forests in this region, this assessment should be supplemented with additional resources and stand-level information. The high diversity in landforms, microclimates, hydrology, and species assemblages across the assessment area greatly complicates model projections and interpretation. In this assessment, forest ecosystems were combined and generalized based on NatureServe’s ecological systems, which are themselves made up of hundreds of unique “associations” (Chapter 1). Forest ecosystems have the potential to manifest themselves in very different ways across the assessment area (e.g., varying in species associations and landscape position), and it is important to have a good working knowledge of forest ecosystems at the local level in each section. It is essential to consider local characteristics such as past management history, soils, topographic features, species composition, forest health issues, and recent disturbances when interpreting these general vulnerabilities at local scales. Some site-level factors may amplify these expected vulnerabilities, yet others may buffer the effects of climate change. Developing a clear understanding of potential vulnerabilities across relevant scales will then enable forest managers, landowners, planners, and other resource specialists to consider appropriate adaptation responses. This is true whether the task is to manage a single stand over a few years, or to design a long-term management plan for a large tract of land. In the following chapter, we extend the discussion to consider the implications of climate trends and forest ecosystem vulnerabilities for other ecosystem services and resource areas that are often important to forest managers. ChAPTER 7: MANAGEMENT iMPLiCATioNS The previous chapters of this assessment have described observed and anticipated climate trends, potential impacts to forest ecosystems, and the climate-related vulnerability of nine forest ecosystems in the assessment area. This chapter takes one additional step and summarizes some implications of these climate change impacts and vulnerabilities for a variety of topics important to forest managers. Changes in climate, impacts on forest ecosystems, and ecosystem vulnerability will combine to create both challenges and opportunities in forest management. Topics were selected to encompass major resource areas that are priorities for public and private land managers. These topics, and the descriptions of climate change implications, are not comprehensive. Some topics have received less scientific attention or contain greater uncertainty. For some topics we relied on input from subject-area experts to discuss climate change implications. Our goal is to provide a springboard for thinking about management implications of climate change and to connect managers to other relevant resources. When available, the “more information” sections provide links to key resources for managers to find more information about the impacts of climate change on that particular topic. The topics addressed are: wildlife, threatened and endangered plant species, nonnative invasive plant species, fire and fuels, infrastructure, air and water quality, forest products, nontimber forest products, forest carbon, recreation, wilderness, cultural resources, urban forests, forest-associated towns and cities, and planning for conservation and natural resource management. This chapter does not make recommendations as to how management should be adjusted to cope with climate impacts. We recognize that the implications of climate change will vary by ecosystem, ownership, and management objective. Therefore, we provide broad summaries rather than focusing on particular management issues. A separate document, Forest Adaptation Resources, has been developed to assist land managers in a decisionmaking process to adapt their land management to projected impacts (Swanston and Janowiak 2012). WILDLIFE Climate change is likely to have both shortand long-term effects on individual organisms, populations, species, and wildlife communities in the Central Appalachians region. These effects may range from direct habitat loss to complex indirect impacts on wildlife populations and their habitats. Changes to habitats discussed in Chapter  will likely result in range expansion for some species and the reduction or complete loss of available suitable habitat for others. Wildlife populations may respond by adapting to new conditions or migrating to follow shifts in suitable habitat; species that are unable to adapt or have limited dispersal ability, particularly those that are already rare, may face substantial challenges in a changing climate. Managing wildlife species may require adjustments to accommodate shifting ranges or to provide supplemental food sources during critical periods. Climate change vulnerability assessments have been conducted for many individual species within West Virginia (Byers and Norris 2011), and the broader Central 19 ChAPTER 7: MANAGEMENT iMPLiCATioNS Appalachians region (Furedi et al. 2011, Schlesinger et al. 2011). These assessments are generally focused on state-listed sensitive species for which climate change is only one of a multitude of stressors which have already affected population ranges or viability. The Climate Change Bird Atlas uses forest inventory data and species-specific habitat requirements to examine the potential for climate change to alter the distribution of 147 bird species across the eastern United States (Landscape Change Research Group 2014). Birds appear to be less vulnerable to climate change impacts than other taxonomic groups because they tend to have less habitat specificity, are able to disperse long distances, and are not as hindered by natural and anthropogenic obstacles on the landscape. However, bird species that are dependent on specific habitat types (e.g., high-elevation conifer forest) may be unable to meet their habitat requirements in a new location, or habitat shifts may introduce new competitors and predators (Matthews et al. 2011a). Other potential climate change impacts include changes in the timing of migration for some birds, or the resources (e.g., flowers, seeds, larvae) upon which they depend. Many short-distance migrants have been observed to respond to local changes by adjusting their arrival or departure dates, but long-distance (e.g., transcontinental) migrants respond to cues at their origin, and are unable to predict conditions at their summer grounds (Hurlbert and Liang 2012). Birds arriving either too early or too late could face suboptimal conditions (e.g., limited food resources or difficulty finding mates), resulting in adverse impacts to fitness and survival (Fraser et al. 2013). A young bird amid rhododendron and hemlock. Photo by Patricia Butler, Northern Insitute of Applied Climate Science (NIACS) and Michigan Tech, used with permission. 170 ChAPTER 7: MANAGEMENT iMPLiCATioNS Bat species that rely on insects for food after emerging from hibernation may face similar challenges; shifts in insect populations can influence bats’ ability to regain weight lost during hibernation and to reproduce successfully. Bats may be particularly sensitive to climate change because many aspects of their ecology and life history are closely tied to temperature and precipitation, and many species in the assessment area have already suffered catastrophic declines as a result of whitenose syndrome. Modeling of Indiana bat habitat used maternity habitat requirements of less than 82 °F and projected the summer range to contract to climatic refugia in the northeastern United States and Appalachian Mountains (Loeb and Winters 2013). Other mobile mammal species found in the assessment area may face similar range reductions, particularly species that are adapted to cool, moist habitats. However, species that are dependent on a narrow range of conditions or have limited mobility may not be able to shift to alternate locations as climate and habitat conditions change. For example, the West Virginia northern flying squirrel is closely tied to high-elevation spruce/northern hardwood forests and is restricted in its ability to exploit alternative habitats because of competition with the southern flying squirrel. In addition, as the habitat and range of the southern flying squirrel expands in response to climate change, the potential for hybridization (and loss of genetic integrity) with northern flying squirrels increases (Garroway et al. 2010). Squirrels and other wildlife species that depend on mast trees may benefit from increases in those tree species projected to do well, such as post and white oaks and pignut hickory. Most regional amphibians and fish are poor dispersers and less able to shift to alternate locations in response to adverse changes in local habitat conditions. In addition, many of these species are aquatic or closely associated with specific aquatic and wetland habitats. As a result, these taxonomic groups make up the majority of species considered to be extremely or highly vulnerable in the state vulnerability assessments noted above. The exceptions are cave-obligate species, because caves and associated groundwater-fed aquatic systems appear to be largely buffered from climatic changes. Vulnerable amphibians include the Cheat Mountain and green salamanders, which are constrained by narrow habitat niches; the Jefferson salamander, which is dependent on ephemeral wetlands; and the eastern hellbender and eastern spadefoot toad, which require specific aquatic and riparian habitat features. Mollusk and fish species are threatened by natural and anthropogenic barriers to movement, and physical habitat specificity contributes to their vulnerability to changes in water temperature and precipitation patterns (Byers and Norris 2011). As a group, mollusks are especially vulnerable to negative impacts associated with climate change because of their limited dispersal ability and dependence on a few fish species to serve as larval hosts. Cold- and cool-water fish species, such as brook trout, sculpin species, and redside dace are highly vulnerable to climate change impacts, particularly populations inhabiting small, high-elevation streams that may experience drying of stream beds or elevated water temperatures. Some reptiles and invertebrates are also likely to be affected by climate change. Reptiles rely on ambient environmental temperature to maintain their physiological processes and are uniquely sensitive to changes in temperature. The sex of offspring of many turtle species is determined by ambient temperature; thus, concerns for already sensitive species such as the spotted turtle and bog turtle include physiological impacts that may affect long-term fitness of a population regardless of vegetative habitat changes. Although some research has been conducted on how climate change might affect insects, most of that work is focused on European butterflies and insects of economic and environmental concern in forestry and agriculture 171 ChAPTER 7: MANAGEMENT iMPLiCATioNS Eastern garter snake. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. (Andrew et al. 2013). Within the assessment area, climate-related changes in hydrology and declines in stream quality are expected to adversely affect several dragonflies, such as the crimson-ringed whiteface, rapids clubtail, and green-faced clubtail (Furedi et al. 2011). The topic of climate change impacts to fish and wildlife is an area of very active research, with new insights into species’ adaptations and management ideas to help populations meet these challenges emerging constantly. In addition to research publications, several tools are available online to assist land managers in evaluating species vulnerabilities and potential changes to fish and wildlife resources. A few of these resources follow: 172 More Informaion • The U.S. Forest Service’s Climate Change Resource Center provides information related to climate change impacts to wildlife and species’ responses: www.fs.fed.us/ccrc/topics/wildlife/. Please note that Web addresses are current as of the publication date of this assessment but are subject to change. • NatureServe’s Climate Change Vulnerability Index (CCVI) tool uses readily available information about a species’ natural history and distribution and about the landscape to predict whether it will likely suffer a range contraction and population reductions due to climate change: https://connect.natureserve.org/science/climatechange/ccvi ChAPTER 7: MANAGEMENT iMPLiCATioNS • The Climate Change Bird Atlas is a companion to the Climate Change Tree Atlas and uses information about the direct climate effects as well as changes in habitat to project changes in bird species distributions: www.nrs.fs.fed.us/atlas/bird/ • The Appalachian Landscape Conservation Cooperative (LCC) Web site provides links to resources, documents, papers, webinar series announcements, and other information about drivers and impacts of climate change (including those affecting wildlife and fish), particularly in relation to the Appalachian landscape: http://applcc.org/resources/climate-change • Many states are working to incorporate climate change information into their state wildlife action plans. Voluntary guidance has been provided by the Association of Fish and Wildlife Agencies: www.fishwildlife.org/files/AFWA-VoluntaryGuidance-Incorporating-Climate-Change_ SWAP.pdf • West Virginia’s Climate Change Vulnerability Assessment for Species of Concern provides evaluations of climate change impacts for many plants and animals in the assessment area based on NatureServe’s CCVI: http://wvdnr.gov/publications/PDFFiles/ ClimateChangeVulnerability.pdf • The Ohio Department of Natural Resources (ODNR) provides a Web page with a variety of links to vulnerability assessment resources: http://www.dnr.state.oh.us/ Home/ExperienceWildlifeSubHomePage/ where_to_viewwildlifelandingpage/ OldWomanCreekDefault/ClimateandWildlife/ climate_wlvulnerability/tabid/2372/Default.aspx ThREATENED AND ENDANGERED PLANT SPECiES The Central Appalachians region contains a great diversity of threatened, endangered, and rare plants. Within the assessment area, the U.S. Fish and Wildlife Service (USFWS) lists eight plant species as threatened or endangered (T&E): running buffalo clover, northern wild monkshood, eastern prairie fringed orchid, Virginia spiraea, small whorled pogonia, northeastern bulrush, harperella, and shale barren rockcress (USFWS 2014). These species occur in habitats that include wetlands, riparian areas, deciduous forests, grasslands, and small patch habitats such as shale barrens. In addition, the U.S. Forest Service lists 81 plant species as Regional Forester’s Sensitive Species due to their rarity on the Monongahela and Wayne National Forests. State Natural Heritage Programs track many more rare plant species, including well over 400 species in West Virginia alone, with additional species tracked in the Appalachian portions of Ohio and Maryland. Thus, rare plants can be found in all of the forest ecosystems that are included in this assessment. Given the numerous habitats in which rare plants are found, the effects of climate change on rare plants are likely to vary widely. In general, species with limited distributions are believed to be disproportionately vulnerable to the negative impacts of climate change because suitable habitat may not be available, or because they have no way of migrating to suitable habitat that may become available (Schwartz et al. 200b). However, predicting impacts on individual species can be difficult because many rare species may be limited by narrow ecological tolerances that are not related to climate sensitivity (Schwartz et al. 200b). 173 ChAPTER 7: MANAGEMENT iMPLiCATioNS The West Virginia Natural Heritage Program applied NatureServe’s Climate Change Vulnerability Index to 18 rare plant species and predicted that 7 of them would be highly vulnerable or extremely vulnerable to negative impacts, including 4 T&E species (northeastern bulrush, harperella, small whorled pogonia, and shale barren rockcress). Eight rare species were predicted to be moderately vulnerable, including two T&E species (Virginia spiraea and running buffalo clover). Only four (non-T&E) rare plant species were predicted to remain stable under a changing climate (Bentley’s coralroot, Torrey’s mountainmint, Tennessee pondweed, and lillydale onion). Increased fire may benefit some threatened and endangered plants by maintaining habitat or promoting flowering, as evidence suggests for running buffalo clover and eastern prairie fringed orchid (Hessl and Spackman 1995). More Informaion • U.S. Fish and Wildlife Service Endangered Species Database: http://www.fws.gov/endangered/ • NatureServe’s Climate Change Vulnerability Index: https://connect.natureserve.org/science/ climate-change/ccvi • Ohio Natural Heritage Database and Ohio Rare Plant List: http://www.dnr.state. oh.us/Home/wild_resourcessubhomepage/ ResearchandSurveys/OhioBiodiversityDatabase/ tabid/2352/Default.aspx • Maryland Natural Heritage Program Rare, Threatened, and Endangered Plants: http://dnr.maryland.gov/wildlife/Plants_Wildlife/ rte/rteplants.asp • West Virginia Natural Heritage Program Rare, Threatened, and Endangered Species: http://www.wvdnr.gov/Wildlife/Endangered.shtm • Climate Change Vulnerability Assessment of Species of Concern in West Virginia (Byers and Norris 2011): http://wvdnr.gov/publications/ PDFFiles/ClimateChangeVulnerability.pdf 174 Wildlowers of West Virginia. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. NoNNATiVE iNVASiVE PLANT SPECiES Various researchers and predictive models suggest that climate change will likely increase the ability of many invasive plants to invade and spread (Alpert et al. 2000, Dukes et al. 2009, Hellmann et al. 2008). However, the overall impact of invasive plants will vary based on individual species responses, and in some cases the distributions of invasive plants may decrease (Bradley et al. 2009). In general, increased invasions of warm climate species and decreased invasions of cold climate species might be expected in the assessment area. Projected increases in fire activity and disturbances related to extreme weather may favor the expansion of disturbance-adapted invasive species, especially southern climate species like cogongrass and kudzu. Cogongrass in the southeastern United States has contributed to altered fire regimes and is expected to advance northward with warmer temperatures (Lippincott 2000). In addition, a changing climate has the potential to affect the life cycle of invasive species that are already established in the assessment area. The phenology of temperate plants, such as flowering and leaf-out dates, has been well documented and ChAPTER 7: MANAGEMENT iMPLiCATioNS is known to be especially sensitive to temperature (Cleland et al. 2007b, Parmesan and Yohe 2003). Species that are most responsive to temperature in terms of their flowering date—that is, species that flower earlier in warm years and later in cold years—are the ones that will likely increase in abundance in the face of climate change. Research has shown that many nonnative invasive plants have more flexible flowering dates and have shifted these dates to earlier in the spring than native plants or even nonnative plants that are not invasive (Primack and Miller-Rushing 2012, Willis et al. 2010). For example, the invasive plant purple loosestrife was found to bloom several weeks earlier than it did a century ago, whereas the flowering dates of many other species, such as most native lilies and orchids, did not shift (Primack and Miller-Rushing 2012). As invasive plant invasions become more widespread, forest managers may need to invest more resources to control invasive plant populations and minimize impacts to forests (e.g., prescribed burns and timber harvest). For example, ailanthus is a particularly problematic invasive species that has already been increasing in the assessment area and may benefit from climate change. Data from the Wayne National Forest show ailanthus trees >5 inches in diameter have increased from 0.7 percent of cover to 1. percent in a little over a decade; without action this species is likely to increase exponentially. The Wayne National Forest is working with ODNR, the Appalachian Ohio Weed Control Partnership, and the U.S. Forest Service’s Northern Research Station and Northeastern Area State & Private Forestry to aerially map and strategically treat 500,000 acres of ailanthus across all ownerships in southeastern Ohio. These collaborators are able to identify pockets of heavy infestations that can be treated with standard herbicide treatments and future experimental control with a biological agent. More Informaion • Appalachian Ohio Weed Control Partnership: http://appalachianohioweeds.org/ • Huebner and partners at the USFS Northern Research Station, West Virginia University, and Ohio State University are currently finishing a 4-year study that looks at the impacts of timber harvesting and prescribed fire on three invasive species (garlic mustard, Japanese stiltgrass, and ailanthus) in comparison to red oak. This study may shed more light on the likelihood of increased invasions due to climate change-related increases in disturbance: Huebner, C.D.; McGill, D.; Matlock, G.; Minocha, R.; Dickinson, M.; Miller, G. (unpublished work). Defining an effective forest management strategy that deters invasion by exotic plants: invasive plant response to five forest management regimes. For more information, visit http://nrs.fs.fed.us/people/ chuebner. FiRE AND FuELS Potential climate change impacts include an increase in wildfire risk, especially during summer and fall. As mentioned in Chapter 5, invasive shrub and herbaceous cover can increase fuel abundance, as can mortality of native plants. Increased levels of downed woody debris resulting from winter storm and wind events can also contribute to dry fuel loads. There are three fire seasons in the assessment area: spring, late summer, and fall. The spring season generally lasts from March through late April before leaf-out, and provides the longest burn window when fuels are dry. By the end of May or early June, green-up of understory vegetation raises fuel moisture and tree leaf-out prevents adequate daytime drying of fuels. The late summer season generally lasts from late August through September, 175 ChAPTER 7: MANAGEMENT iMPLiCATioNS when leaves and ground fuels begin to dry out. Droughty weather or at least 7 to 10 rain-free days are necessary for fuel moisture to be low enough to burn. The fall season typically begins in mid- to late October after the first hard frost (and the start of leaf fall) and runs through November. This burn window is extremely variable and fire behavior can be more extreme in fall due to the presence of dry leaf litter, especially oak, that has not yet been compressed by rain or snow. Projected changes in climate could affect the ability to apply prescribed fire in the assessment area. In spring, increased rainfall could make it difficult to conduct prescribed burns. Throughout the spring and summer, changing precipitation patterns, such as intense rain events followed by longer dry periods, could result in longer periods of drier burn conditions. Burning under drier conditions may result in more intense and hotter fires, including fires that use ladder fuels to move into the forest canopy. As the growing season is extended later into the fall, there is even more potential for increased fuels accumulation. On an interannual level, drought increases wildfire risk during all fire seasons (Lafon et al. 2005) and is likely to play a critical role in future shifts in fire windows and behavior. Shifts in climate that result in a longer fire season or extension of critical fire weather days would, in turn, increase the potential risk of wildland fire. Change in fire risk across the assessment area and its impacts at local scales will depend on both land use and management decisions. Potential management responses might include rescheduling prescribed burns as optimal burn windows shift toward summer and fall. Fuel models may also need to adjust to climate-related vegetation changes such as increased density of invasive plants, or shifts in species composition that affect fuels on the forest floor (e.g., from maple to oak). Policy and funding decisions and public attitude will ultimately define the 17 response that makes the most sense, and responses may differ between landowners, land managers, and organizations. More Informaion • The U.S. Forest Service’s Climate Change Resource Center: Wildfire and Climate Change: www.fs.fed.us\ccrc\topics\wildfire\ • The Consortium of Appalachian Fire Managers and Scientists (CAFMS): www.cafms.org iNFRASTRuCTuRE Many landowners and agencies are responsible for managing infrastructure on the forested landscape, such as roads, power lines, sewer lines, dams, drainage ditches, and culverts. Specifications for water infrastructure are based on past climate patterns, and the current trend of intensifying precipitation has placed additional strain on outdated infrastructure. Storms, extreme temperatures, longer growing seasons, and warmer winters can pose particular challenges for infrastructure. Extreme heat and longer growing seasons can result in rising costs associated with roadside and power line vegetation management. Extreme cold and freeze-thaw cycles can accelerate road deterioration. Intense rainfall could increase the potential for erosion on dirt and gravel roads common in forest landscapes, logging projects, gas development, and rural areas. Water resource infrastructure such as bridges, sewers, major culverts, low-water crossings, and dams may have to be redesigned and rebuilt to accommodate flows of increased duration and intensity. Improved stream bank stabilization may have to be incorporated to prevent scouring. Costs associated with debris removal in waterways could also rise. Projected increases in average temperature, summer heat waves, and summer storms are expected to place additional strain on electrical infrastructure. ChAPTER 7: MANAGEMENT iMPLiCATioNS An old culvert. Land managers are beginning to replace culverts like this one with larger culverts designed to accommodate larger peak lows and allow the passage of aquaic organisms. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. Although not directly attributed to climate change, an extreme weather event can serve to illustrate the impacts of such events on electrical systems. On June 29, 2012, a derecho with sustained winds of 0 miles per hour gusting to 100 miles per hour ravaged a 00-mile swath across 11 states including Ohio, West Virginia, and Maryland. Across the region, 4.2 million electrical customers lost service. West Virginia, a rural state with sparse populations and mountainous topography, was particularly devastated; more than 00,000 customers lost power for 10 days or more (U.S. Department of Energy 2012). According to an analysis by the U.S. Department of Energy, extensive debris, downed-tree removal operations, additional storms, and unusually high heat hindered the restoration of power (U.S. Department of Energy 2012). The derecho made the National Oceanic and Atmospheric Administration’s (NOAA’s) list of billion-dollar weather events ($2.8 billion) and resulted in the death of 28 people (NOAA 2014a). Dominion Power reported the derecho to be the most severe weather event in the company’s 100-year history after Hurricanes Irene and Isabel (Knight 2012). Following the derecho, the region experienced record high temperatures, which complicated efforts to restore power. Although millions of residents had to go without air conditioning during this particular storm, heat waves are expected to increase in frequency and duration, and are likely to put great demand on electricity supply. More Informaion • American Society of Civil Engineers 2013 Report Card for America’s Infrastructure: http://www. infrastructurereportcard.org/a/#p/home • The U.S. Energy Sector Vulnerabilities to Climate Change and Extreme Weather: http://energy.gov/ sites/prod/files/2013/07/f2/20130710-EnergySector-Vulnerabilities-Report.pdf 177 ChAPTER 7: MANAGEMENT iMPLiCATioNS AIR QUALITY The direct and indirect effects of a changing climate have important implications for air quality and its management. Although future changes in pollutant emissions can be estimated, air quality impacts will continue to be strongly influenced by climate and weather variables, such as temperature, humidity, and air flow (Mickley et al. 2004). Because mercury, nitrogen, and sulfur are deposited onto the landscape through rain and snow, projected increases in precipitation may increase atmospheric deposition, thus increasing mercury contamination and the acidification of soils and surface waters (Driscoll et al. 2007). Tropospheric ozone in the Central Appalachians is projected to increase as a result of higher temperatures and decreased ventilation resulting from changes in air flow (Wu et al. 2008). Because heat waves and air stagnation retain ozone levels for extended periods, these climate changes affect ozone pollution episodes more than mean ozone levels, and are projected to offset and surpass decreases in ozone brought about by regulation (Wu et al. 2008). There is evidence that warmer temperatures and the burning of vegetation can result in increased volatilization of mercury soil reservoirs, potentially releasing mercury into the atmosphere and transferring it between ecosystems, with deposition occurring in a more mobile and toxic form (Jacob and Winner 2009). Particulate matter may also be affected by changes in climate, although changes are less predictable than for ozone. Because particulate matter is cleaned from the air by rainfall, increases in precipitation frequency due to climate change could have a beneficial effect. However, other climate-related changes in stagnant air episodes, wind patterns, emissions from vegetation, wildfire, and the chemistry of atmospheric pollutants will also influence particulate matter levels in different ways. Air quality regulations are important in controlling emissions, but when climate change impacts are taken into consideration, the current thresholds may not be adequate to meet air quality targets. 178 More Informaion • Integrating Knowledge to Inform Mercury Policy: www.mercurynetwork.org.uk/policylinks/ mercury-and-climate-change/ • The Monongahela National Forest monitors wet deposition, dry deposition, ozone, and particulate matter using the National Atmospheric Deposition Program National Trends Network (NADP/NTN), Clean Air Status and Trends Network (CASTNET), and Interagency Monitoring of Protected Visual Environments (IMPROVE): www.fs.usda.gov/Internet/FSE_ DOCUMENTS/fsm9_01135.pdf • Researchers at Penn State are investigating the effects of soil acidification on sugar maple decline on the Allegheny National Forest: http:// ecosystems.psu.edu/directory/wes WATER QUALITY It is widely accepted that streamflow is primarily governed by climate, watershed morphology, and land cover, and that hydrology largely controls sediment and nutrient export. Any change in state variables that alters watershed hydrology also influences water quality dynamics (Likens and Bormann 1995). Climate change is already creating challenges in water management by affecting water availability (Georgakakos et al. 2014). Projected increases in total precipitation in spring, intense precipitation events, and storm frequency are expected to lead to more runoff at that time of year, and a subsequent reduction in water quality arising from increased erosion and sedimentation (Liu et al. 2008, U.S. Environmental Protection Agency [EPA] 1998). Increased runoff also promotes flushing of nutrients (e.g., nitrogen and phosphorus) that build up in natural and disturbed ecosystems, thereby increasing the potential for downstream eutrophication and hypoxia (Peterjohn et al. 199, Vitousek et al. 2010). Additional factors such as fire and insect defoliation exacerbated by climate change are also expected to increase runoff, erosion, and ChAPTER 7: MANAGEMENT iMPLiCATioNS The steep ridges and valleys in the Allegheny Mountains. These landforms are at greater risk of high-velocity runof and erosion. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. sedimentation. Late summer soil moisture deficits combined with a longer growing season have the potential to decrease runoff in the latter half of the year, thereby decreasing the capacity of a stream system to dilute larger loads of nutrients (Delpla et al. 2009). Anthropogenic activities have already damaged aquatic ecosystems by increasing soil erosion and stream sedimentation rates, fragmenting aquatic habitats, reducing channel and floodplain functions, degrading habitats, acidifying and burying streams, and otherwise altering watershed hydrology. Under the range of projected climate changes, aquatic ecosystems would tend to have more varied and more extreme environmental conditions. Changes of this nature tend to place additional hardship on these systems and can further compromise various aquatic resource conditions such as habitat suitability. Aquatic ecosystems that were once intact and naturally functioning can be repaired to various degrees under accelerated timeframes through restoration actions. Accelerating the rate of recovery back toward their inherent state can increase the resiliency of these systems to stressors and disturbances. Water resource managers may minimize risks and impacts by accommodating expanding floodplains, redesigning stormwater and sewer systems, restoring and managing wetlands for stormwater management, and developing novel ways to buffer intense runoff, such as through green roofs and other infrastructure (U.S. EPA 2008). More Informaion • National Climate Assessment: Water Resources: http://nca2014.globalchange.gov/report/sectors/ water • National Water Program 2008 Strategy: Response to Climate Change: http://water.epa.gov/scitech/ climatechange/upload/2008-National-WaterProgram-Strategy-Response-to-Climate-Change. pdf 179 ChAPTER 7: MANAGEMENT iMPLiCATioNS FoREST PRoDuCTS The forest products industry is important to the economies of the assessment area (Chapter 1). Tree species and forest composition are projected to change over the 21st century (Chapters 5 and ). Changes in forest composition across the landscape will be influenced by forest management, and will in turn influence forest management and the forest products industry. Several commercially important species, such as black cherry and sugar maple, are projected to decline significantly under a range of possible climate futures during the next century. Conversely, post oak, white oak, and shortleaf pine are projected to increase in the assessment area. Large potential shifts in commercial species availability may pose risks for the forest products sector if the shifts are rapid and the industry is unprepared. The forest products industry may benefit from awareness of anticipated climate trends and shifts in forest species. In many cases, forest managers can take actions to reduce potential risks associated with climate change or proactively encourage species and forest types anticipated to fare better under future conditions (Swanston and Janowiak 2012). There may be regional differences in forest responses, as well as potential opportunities for new merchantable species to gain suitable habitat in the assessment area. Overall, the effects of climate change on the forest products industry depend not only on ecological responses to the changing climate, but also on socioeconomic factors that will continue to change over the coming century. Major socioeconomic factors include national and regional economic policies, demand for wood products, and competing values for forests (Irland et al. 2001). Large uncertainties are associated with each of these factors. The forest products industry has adjusted to substantial changes over the past 100 years, and continued responsiveness can help the sector remain viable. 180 More Informaion • The U.S. Forest Service 2010 Resources Planning Act Assessment includes future projections for forest products and other resources through the year 200 and examines social, economic, landuse, and climate change influences: www.fs.fed. us/research/rpa/ • The Climate Change Tree Atlas provides information on the projected suitable habitat for tree species under climate change: www.nrs. fs.fed.us/atlas/bird/ NoNTiMBER FoREST PRoDuCTS Hundreds of nontimber forest products are used for food, medicine, craft materials, and other purposes in the assessment area (Chamberlain et al. 2009). Changes in climate will have implications for these products in the assessment area and throughout the broader region. Many of these products will be affected by changes in temperature, hydrology, and species assemblages. As illustrations, effects of climate change on two nontimber forest products with broad cultural and economic importance are discussed briefly here: American ginseng and mushrooms. American ginseng is a perennial herbaceous plant indigenous to the eastern United States that has been traded internationally since the 1700s (Taylor 200). Concerns over the sustainability of wild American ginseng under heavy harvest pressure in Canada, China, and the United States resulted in international protection under the Convention on International Trade in Endangered Species of Wild Fauna and Flora (CITES). The USFWS monitors exports in order to examine trends in wild ginseng harvest and set harvest guidelines and restrictions. Individual states also monitor the harvest and export of ginseng and regulate harvesting. ChAPTER 7: MANAGEMENT iMPLiCATioNS Recent research has examined the potential response of wild ginseng to temperature variations in the assessment area. Findings suggest that genetics of populations at individual locations play a large role in population growth rate responses to temperature. Thus, models that predict responses based on northern and southern boundaries of this species likely underestimate the negative impacts of temperature increases at specific locations (Souther and McGraw 2011). Precipitation may also constrain the overall distribution of ginseng (Souther and McGraw 2011). Neither factor showed a positive population growth response to predicted changes. The combination of harvest pressure and climate change raises concerns about the long-term stability of the American ginseng population in the assessment area. Hunting morels and other mushrooms is a passion for many people throughout the assessment area for their commercial value, medicinal properties, and culinary applications (Emery and Barron 2010). An analysis of fungal fruiting patterns from southern England over a 5-year-period showed a lengthened fruiting period from 33.2 days in the 1950s to 74.8 days in the current decade (Gange et al. 2007). This change corresponded to increased temperatures in August through October. Another study of 83 species in Norway found an average delay in fruiting of nearly 13 days since 1980, coinciding with warming temperatures and a longer growing season (Kauserud et al. 2008). Although longer growing seasons have lengthened the fruiting season of some fungal species, and shifted the timing of fruiting later in the spring and fall, future responses to changes in temperature and precipitation may be tightly linked to local conditions rather than broad geographic trends. Management of nontimber forest products may require increased monitoring of habitats to ensure viable populations under changing conditions. More Informaion • Forest Farming: www.extension.org/forest_ farming • Connecting Non-timber Forest Products Stakeholders to Information and Knowledge: A Case Study of an Intranet Web Site: www.srs. fs.usda.gov/pubs/gtr/gtr_srs11/gtr_srs11-04.pdf • Using Local Ecological Knowledge to Assess Morel Decline in the U.S. Mid-Atlantic Region: www.nrs.fs.fed.us/pubs/jrnl/2010/nrs_2010_ emery_001.pdf FoREST CARBoN Forest carbon sequestration can mitigate greenhouse gas emissions in the atmosphere. However, climate change and indirect impacts to forest ecosystems may change the ability of forests in the Central Appalachians to store carbon. In this assessment, carbon dioxide fertilization effects on forest ecosystems were not directly modeled or assessed, but are considered an important implication for forest management. Within the assessment area, climate change is projected to lead to longer growing seasons and warmer temperatures, which potentially could support increased forest productivity and carbon storage, as long as water and nutrients are available for photosynthesis. This increase could be offset by climate-related disturbances, such as more insect pests or disease, leading to increases in carbon storage in some areas and decreases in others (Hicke et al. 2012, Knicker 2007). Increases in ozone would reduce photosynthesis and carbon sequestration (Felzer et al. 2003). The greatest impacts on forest carbon storage will likely occur through changes in species composition. Habitat suitability models forecast shifts in tree species’ geographic ranges in response to climate 181 ChAPTER 7: MANAGEMENT iMPLiCATioNS changes (Chapter 5). Within the assessment area, the oak/gum/cypress and spruce/fir forests store the most carbon per acre, followed by maple/beech/birch and elm/ash/cottonwood forests (Chapter 1). Oak/hickory and oak/pine forests contain considerably less carbon per acre, but are projected to be more resilient to or even benefit from climate change. Not all forests store carbon in the same pools; for example, oak/hickory forests store more carbon aboveground than in the soil and the spruce/ fir group stores more carbon in the soil. Thus, shifts in species or assemblages of species on the landscape may result in shifts in carbon storage. Invasive plant species also have the potential to alter species composition and ecosystem functioning. The invasive tree ailanthus can increase carbon cycling rates and alter soil chemistry to favor rapid growth and subsequent forest colonization by new ailanthus seedlings (Gómez-Aparicio et al. 2008). Carbon management and conservation of carbon stocks will require managing species composition and maintaining forest cover on the landscape. The biggest loss of forest carbon in the assessment area has already occurred as a result of historic logging, loss of soil from erosion and volatilization from fire, and decades of land conversion from forests to agriculture and urbanization. The ability of existing ecosystems to sequester carbon may be further hindered by increased disturbances and stresses brought on by climate change. Carbon management can benefit future landscape-scale restoration projects. Riparian restoration and wetland restoration have the potential to help landscapes slow the export of nutrients or even capture and store soil carbon that would otherwise leave the watershed. Replanting riparian areas and encouraging the regrowth of these areas can help to address the historic forest carbon loss for several of the ecosystems analyzed in the assessment. Opportunities to focus restoration management on stabilizing soils, planting trees, and addressing historic land degradation are numerous. More Informaion Soil from federal lands. These public lands contain the highest density of carbon in the Central Appalachians. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 182 • The U.S. Forest Service’s Climate Change Resource Center: Forests and Carbon Storage: www.fs.fed.us/ccrc/topics/forests-carbon/ • A Synthesis of the Science on Forests and Carbon for U.S. Forests: www.fs.fed.us/rm/pubs_other/ rmrs_2010_ryan_m002.pdf ChAPTER 7: MANAGEMENT iMPLiCATioNS The Blackwater Falls in the Canaan Valley, West Virginia, a popular recreaion area. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. RECREATioN Opportunities for outdoor recreation depend on the natural resource (e.g., spelunking in caves versus hiking mountain trails) and the weather on any given day. Projected increases in temperature and precipitation, especially heat waves and intense precipitation events (Chapter 4), are expected to change recreation patterns. Warmer spring and fall weather may increase the length of the recreation season, which could require a shift in the open season for recreation areas, requiring more staff hours and potentially more infrastructure. Regional increases in average temperatures and heat waves during summer months could shift visitor behavior, depending on the magnitude of changes. Many visitors to the Monongahela National Forest arrive during the summer to escape the heat at lower elevations or in urban areas, and temperature increases could result in higher visitation rates (Loomis and Crespi 1999, Mendelsohn and Neumann 2004, Richardson and Loomis 2004). If temperatures become too hot for outdoor recreation, however, visitation and outdoor recreation and tourism could decrease (Nicholls 2012). Specific activities such as fishing or skiing may also be limited by warmer temperatures (Morris and Walls 2009). 183 ChAPTER 7: MANAGEMENT iMPLiCATioNS Projected increases in intense precipitation and strong storm events could lead to more frequent closings of public recreation areas. The same derecho of 2012 that knocked out power to several states also blew down thousands of trees across the region, and caused many public places to close, such as Lake Sherwood Recreation Area in West Virginia. In 2013, effects of Hurricane Sandy closed a large part of the Monongahela National Forest for several months because of such hazards as broken, hanging, and down trees, which damaged facilities, and closed roads and trails. Many recreation areas are located near rivers and streams, which are regularly subject to flood events. To properly protect recreation visitors, short- and long-term closings may be needed to repair damage caused by intense precipitation and strong storm events. Warmer winter temperatures could also affect winter recreation. Warmer temperatures that prevent Lake Erie from freezing may allow more moisture to evaporate from the lake and fall as snow on land. However, warmer average temperatures may also increase the probability that precipitation will fall as rain rather than snow. A particular economic concern is the decreased viability of downhill skiing during the holiday season, which can generate as much as one-third of a ski resort’s annual revenue (Dunnington 2011). Although downhill ski areas can generate artificial snow, few options exist for adapting cross-country skiing, sledding, snowshoeing, and other snow-dependent winter sports to warmer temperatures (Morris and Walls 2009). These winter activities may be replaced by hiking and other activities not dependent on snow, requiring adjustments in how recreation areas are managed. The degree of climate change will ultimately influence the severity of impacts on recreation activities, but there are many opportunities for visitors and managers to adapt their activities by changing the timing or location (Morris and Walls 2009). 184 More Informaion • National Climate Assessment Midwest Technical Input Report: Recreation and Tourism Sector: glisa.msu.edu/docs/NCA/MTIT_RecTourism.pdf • Climate Change and Outdoor Recreation Resources: www.rff.org/RFF/Documents/RFFBCK-ORRG_ClimateChange.pdf WILDERNESS The Wilderness Act of 194 was established to protect areas in their natural condition and to assure that an increasing human population, accompanied by expanding settlement and growing mechanization, does not modify all areas within the United States (Wilderness Act of 194). U.S. Forest Service policy directs the agency to “manage the wilderness resource to ensure its character and values are dominant and enduring” (U.S. Forest Service 2007). According to the Monongahela National Forest Land and Management Plan, management emphasis for its eight wilderness areas on the Forest is to “preserve wilderness attributes and the natural environment for future generations” (U.S. Forest Service 200a). It has been argued that climate change would have the greatest impacts on species that are confined to protected areas, largely because populations would not be able to migrate with changing range limits for species (Peters and Darling 1985). Additionally, species within protected areas would potentially face new competitors, predators, or diseases as many native and nonnative species move around on the landscape. Models of climate change impacts on ecosystems project that more than 40 percent of Canada’s protected areas will undergo a major change in vegetation (Lemieux and Scott 2005). Management of wilderness areas may need to address difficult questions about whether to protect current species assemblages, or to allow new species assemblages to form, and if the latter, to what extent. ChAPTER 7: MANAGEMENT iMPLiCATioNS An increase in intense precipitation and strong storm events would cause muddy conditions on trails, erosion of trail tread, and down trees across trails. Mechanized equipment is not allowed in wilderness areas; the additional physical labor to complete trail maintenance is expensive and time consuming. For example, after Hurricane Sandy in 2013, trails within Otter Creek Wilderness and Cranberry Wilderness areas were closed for several months until specialized crews were funded to clear the trails with crosscut saws and axes. Responding to increased disturbances may require additional resources to manage wilderness areas. More Informaion • Climate Change Toolbox: Effect of Climate Change on Wilderness and Protected Areas: www.wilderness.net/climate • The U.S. Forest Service’s Climate Change Resource Center: Wilderness and Climate Change: www.fs.fed.us/ccrc/topics/wilderness/ Private lands juxtaposed with the Shawnee State Forest, Ohio. Photo by the Ohio Department of Natural Resources, used with permission. 185 ChAPTER 7: MANAGEMENT iMPLiCATioNS CuLTuRAL RESouRCES The remnants of past human activity, such as paintings, sculptures, and objects for everyday life, are present within the assessment area. These resources date to both prehistoric and historic time periods, and exist both above and below the ground surface. Climate change impacts on the physical environment have the potential to affect the nature, character, and condition of these cultural resources. Increases in extreme precipitation events, in combination with a more episodic regime, are expected to intensify erosion and weathering of cultural resources. Consequently, the physical integrity of historic structures could be undermined and subsurface resources threatened if the soil covering them is washed away. As precipitation increases, the risk of flooding also escalates; flooding would hasten the erosion process of sites on ridge tops and on flood terraces. Floodwaters can further threaten the integrity of historic structures in low-lying areas by eroding the foundation, or adding moisture. The increased moisture can promote mold and fungus growth, thereby hastening deterioration of wooden and other constructed features (Schiffer 199). Erosion of rock shelters has already been witnessed within the assessment area on sites composed largely of erodible sandstone that are more frequently being inundated with water. Artifacts and other cultural materials located in these shelters have been transported by water to nearby creeks. Increased moisture levels and damage from freeze/thaw cycles and subsequent erosion have resulted in roof collapse within these rock shelters as well. Projected increases in freeze/thaw events and deep soil frost would exacerbate these effects. Longer growing seasons and range shifts in native and invasive plants expand the potential for these taxa to damage historic structures as these plants tend to cling to structures at points of weakness, 18 accelerating structural degradation (Schiffer 199). An altered fire regime could become an increasing source of disturbance if climate shifts encourage more frequent or intense fire behavior. Fire and firefighting activities can destroy historic structures and threaten all types of cultural resources (Buenger 2003). Managing cultural resources will become more challenging as a result of the direct and indirect impacts of climate change. Identifying and documenting existing cultural resources now will be critical in conserving these important artifacts and historical information. More Informaion • Climate Change and World Heritage: whc.unesco. org/documents/publi_wh_papers_22_en.pdf • National Park Service Climate Change Response Strategy: http://www.nps.gov/orgs/ccrp/upload/ NPS_CCRS.pdf uRBAN FoRESTS Climate change will likely affect urban forests in the assessment area as well. Urban environments can pose additional stresses to trees not encountered in natural environments, such as pollution from vehicle exhaust, confined root environments, and road salts. Urban environments also cause a “heat island effect,” and thus warming in cities will likely be even greater than that experienced in natural communities. Impervious surfaces can make urban environments more susceptible to flash floods, placing flood-intolerant species at risk. All of these abiotic stressors can make urban forests more susceptible to nonnative species invasion, and insect and pathogen attack, especially because only a limited range of species and genotypes is typically planted in urban areas. Urban settings are also the most likely places for exotic insect pests to be introduced. ChAPTER 7: MANAGEMENT iMPLiCATioNS Projected changes in climate can pose both challenges and opportunities for the management of urban forests. Shifts in temperature and changes in extreme events may have effects on species selection for planting. Native species projected to decline under climate change will likely not tolerate even more extreme conditions presented by urban settings. Conversely, urban environments may favor heat-tolerant or drought-tolerant native species or new migrants (Chapter 5). Determining appropriate species for planting may be a challenge, but community foresters are already familiar with the practice of planting species novel to an area. Because of urban effects on climate, many community forests already contain species that are from planting zones south of the area or cultivars that tolerate a wide range of climate conditions. Large disturbance events may also become more frequent or intense in the future, necessitating informed decisions in response. For example, wind events or pest outbreaks may be more damaging to already stressed trees. If leaf-out dates advance earlier in the spring due to climate change, community forests may be increasingly susceptible to early-season frosts or snowstorms. More people and larger budgets may be required to handle an increase in the frequency or intensity of these events, which may become more difficult in the face of reduced municipal budgets and staffing. More Informaion • The U.S. Forest Service’s Climate Change Resource Center: Urban Forests and Climate Change: www.fs.fed.us/ccrc/topics/urban-forests/ • Urban Forests: Climate Adaptation Guide: www. toolkit.bc.ca/Resource/Urban-Forests-ClimateAdaptation-Guide • Climate Change Adaptation Options for Toronto’s Urban Forest: www.cleanairpartnership.org/pdf/ climate_change_adaptation.pdf FoREST-ASSoCiATED TOWNS AND CITIES The forests of the Central Appalachians are deeply and intimately linked to human communities. Conversely, these communities are tied to the health and functioning of surrounding forests, whether for economic, cultural, recreational, or other reasons. Climate change impacts on forest ecosystems are likely to affect the human communities that use these resources and to change or challenge how those communities use and relate to these forests. These complex feedbacks could very well pose a challenge to current forest management goals and activities. Consequently, it is important to address potential climate change impacts on forest-associated towns, cities, and other communities, and the implications for managing healthy ecosystems. Although impact models can predict species or community responses to climate change, considerably less is known about the potential social and cultural impacts of climate or forest change and how human communities might best respond. Community vulnerability to climate change is a function of the community’s exposure to change, such as being situated within a flood plain projected to receive increased precipitation, and its relative sensitivity to such changes, such as being constrained by reduced funding from the Federal Emergency Management Act (FEMA) due to budget cuts or other national priorities. Community adaptive capacity is a function of the community’s ability to act in an adaptive way and includes both material (i.e., capital) and nonmaterial (i.e., leadership) resources that can be leveraged by the community to monitor, anticipate, and proactively manage hazards, stressors, and disturbances. These concepts help frame the issue of climate change from a community perspective, but it is important to keep in mind that every forest- 187 ChAPTER 7: MANAGEMENT iMPLiCATioNS Muliple land uses in West Virginia. Agriculture and development dominate the lat valleys. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. associated community has particular conditions, capacities, and constraints that might make it more vulnerable or resilient to climate change than other communities. For example, forest users from a city like Huntington, WV, face different sources of vulnerability than forest users from a small town like Glouster, OH. Moreover, the effects of climate change and forest impacts are not evenly distributed geographically or socially. For example, a tourismdependent community may be more or less exposed to climate change than certain social groups 188 within communities (e.g., individuals working in forest products industries), or they may be equally exposed, but more or less able to adapt. If resource professionals, community leaders, and local organizations are to help communities mitigate the impacts of climate change and adapt, they must be able to assess community vulnerabilities and capacities to organize and engage various resources (Fischer et al. 2013). ChAPTER 7: MANAGEMENT iMPLiCATioNS More Informaion • Assessing Resilience in Social-Ecological Systems: Workbook for Practitioners: www. resalliance.org/index.php/resilience_assessment • Assessing Social Vulnerability to Climate Change in Human Communities near Public Forests and Grasslands: A Framework for Resource Managers and Planners: http://people.oregonstate.edu/ ~hammerr/SVI/Fischer_etal_JoF_2013.pdf • Community Vulnerability and Adaptive Capacity Project: http://www.cfc.umt.edu/VAC/default.php • A study is underway to explore the perceived vulnerability and adaptive capacity of forestassociated human communities in southeastern Ohio. For more information, contact Dr. Daniel Murphy at the University of Cincinnati: http:// asweb.artsci.uc.edu/collegedepts/anthro/fac_staff/ profile_details.aspx?ePID=MzA0ODcx CoNSERVATioN PLANNiNG Climate change has many important implications for land conservation planning in the Central Appalachians. Climate change science can be used to help prioritize land conservation investments and help guide project design. Some of the most useful decision-support tools for conservation planning are site-specific technical assistance through scientific experts to geographic information systems (GIS) mapping tools that allow the user to assess how individual parcels of land relate to variables such as forest carbon and projected “climate-safe” habitat areas. Conservation in the complex landscapes of the Central Appalachians also requires careful analysis to evaluate the potential contribution of conservation projects to climate adaptation. The region’s forests provide vital ecosystem services to human and natural communities. These services, such as drinking water supplies and cold-water habitats, could be affected by greater extremes of precipitation and other manifestations of climate change. Given the steep slopes in the region, watersheds are naturally prone to flooding and at particular risk from increases in extreme precipitation events. Conservation linked with adaptive management can be directed to the most vulnerable watersheds to help them withstand these impacts. Further, climate change analysis is nuanced by the region’s globally significant mixture of microhabitats and connecting habitat corridors stretched across rugged landscapes. Planning for conservation of terrestrial habitat “strongholds” from climate change requires a close look at the landscape to identify those corridors and habitats that will be most resilient in the face of projected shifts. As evidence of the unique opportunities in the upper Potomac watershed, the Open Space Institute and Doris Duke Charitable Foundation have targeted a special conservation funding source to this region for conservation of important sites for climate adaptation. The Nature Conservancy’s resilience analysis project identifies sites across the Northeast that have high or low resilience to climate changes based on geophysical characteristics (Anderson et al. 2012). Integrating this kind of information into conservation planning and prioritization can help identify and protect areas that have unique potential for conservation. Carbon dioxide emissions have directly contributed to ongoing climate change, and it is unclear how emissions levels may change over the course of the century. Identifying forest tracts that have high carbon stocks or potential for high carbon levels through conservation- and carbon-oriented management can help maintain and even increase this important source of carbon mitigation. U.S. forests currently sequester 10 to 20 percent of the nation’s carbon emissions each year (Ryan et al. 189 ChAPTER 7: MANAGEMENT iMPLiCATioNS 2010). Carbon-oriented prioritization is particularly important in the Central Appalachians region, where the region’s forests have substantial carbon stores. For example, oak/hickory forests in the region can hold as much as 132 tons of carbon per acre in soils and aboveground biomass. Land managers can prioritize protection on sites that are strong carbon sinks, or that have potential for resilience under climate change. Designing land conservation projects for climate objectives may require specific long-term ownership and management prescriptions to be attached to a conservation agreement. In some cases, a good conservation strategy may be to leave lands in private ownership, and to develop conservation easement terms that support adaptive management by the landowner to address climate shifts. In other cases, where complex restoration or species-specific management is needed, an appropriate conservation strategy might be to seek a public agency owner that can provide the necessary financial and technical resources. In either instance, the key principle is to use available climate information to assess projected stressors on the property in the future, and then to integrate those considerations into project design. All of the efforts described above will be advanced by new science and data products to guide project selection and design. Private nonprofit organizations, government agencies, landowners, and potential funders will increasingly need spatially explicit information on how climate shifts will play out over the land. This science can enable effective use of funding, staff time, and other resources that are essential to advancing “climate-informed” conservation of forests in the Central Appalachians, and shaping conservation efforts to deliver a more resilient landscape. More Informaion • The Open Space Institute and Doris Duke Charitable Foundation: www.osiny.org/site/PageS erver?pagename=Issues_Habitat 190 • The Nature Conservancy Northeast Resilience Analysis: www.conservationgateway.org/ ConservationByGeography/NorthAmerica/ UnitedStates/edc/reportsdata/terrestrial/resilience/ ne/Pages/default.aspx NATuRAL RESouRCE MANAGEMENT PLANNiNG Until recently, climate change has not played a large role in natural resource planning. However, many federal and state-level land management agencies are beginning to address the issue. For example, the U.S. Forest Service’s 2012 Planning Rule directly addresses the impacts and ramifications of climate change. In fact, climate change was among the stated purposes for revising the rule. Similarly, the state forestry agencies of Ohio, West Virginia, and Maryland began to officially address climate change in the 2010 State Forest Resource Assessment and Strategy. Private lands make up about 85 percent of forest lands in the Central Appalachians region (Chapter 1). Northeastern Area State & Private Forestry oversees the Forest Stewardship Program to assist private landowners with conservation planning and to provide forest management plans at low cost. This unit is currently funding two examples of forest adaptation to climate change, using the tools in Forest Adaptation Resources (Swanston and Janowiak 2012) to identify adaptation actions in Forest Stewardship Plans. Because the goals for private landowners are diverse and can include goals for soil and water conservation, timber production, wildlife, and many more values, each example of adaptation will differ based on landowner needs. The Northeastern Area unit is also working with the Northern Institute of Applied Climate Science to develop an online version of the adaptation workbook presented in Forest Adaptation Resources that will be more accessible to natural resource managers. ChAPTER 7: MANAGEMENT iMPLiCATioNS Management plans for national forests or state agencies are typically written to guide management for a 10- to 15-year period, and it may be difficult to foresee projected shifts in climate within this short planning horizon. If climate change results in more frequent disturbances or unanticipated interactions among major stressors, managers may find it more difficult to adhere to the stated goals, objectives, and priorities in current Forest Plans. Incorporating adaptive management principles and including flexibility to address shifting conditions and priorities may be a strategy to handle the uncertainties of climate change. But building that flexibility into forest plans may pose a challenge both in completing the analysis (with specialists who may be unaccustomed to analyzing adaptive management strategies) and in educating the public about the need for proposed actions. Project-level planning on national forests will face challenges with interdisciplinary teams grappling to understand both the impacts that projects may have on greenhouse gas emissions and carbon sequestration levels and the impacts that climate change may have on projects. Input from the public is expected to increasingly question these relationships and interdisciplinary teams must be able to respond. Draft guidance is available from the Council on Environmental Quality (CEQ) and the U.S. Forest Service on project-level climate change considerations. More Informaion • Forest Steward Program for private landowners: http://www.na.fs.fed.us/stewardship/index.shtm • Region 9 Climate Change Guidance: www.fs.fed. us/emc/nepa/climate_change/includes/cc_nepa_ guidance.pdf • Statewide Forest Action Plans: http://www. forestactionplans.org/regions/northeastern-region • Forest Adaptation Resources: Climate Change Tools and Approaches for Land Managers provides concepts and tools for integrating climate change considerations into natural resource planning and management: www.nrs.fs.fed.us/pubs/40543 ChAPTER SuMMARY The breadth of the topics above highlights the wide range of effects that climate change may have on forest management in the Central Appalachians region. It is not the role of this assessment to identify adaptation actions that should be taken to address these climate-related risks and vulnerabilities, nor would it be feasible to prescribe suitable responses for all future circumstances. Decisions to address climate-related risks for forest ecosystems will be affected by economic, political, ecological, and societal factors. These factors will be specific to each land owner and agency, and are highly unpredictable. Confronting the challenge of climate change presents opportunities for managers and other decision-makers to plan ahead, build resilient landscapes, and ensure that the benefits that forests provide are sustained into the future. Resources are available to help forest managers and planners incorporate climate change considerations into existing decisionmaking processes (Swanston and Janowiak 2012) (www.forestadaptation.org). This assessment will be a useful foundation for land managers in that process, to be further enriched by local knowledge and site-specific information. • Draft National Environmental Policy Act Guidance on Consideration of the Effects of Climate Change and Greenhouse Gas Emissions: http://energy.gov/sites/prod/files/CEQ_Draft_ Guidance-ClimateChangeandGHGemissions2.18.10.pdf 191 GLoSSARY aerosol biomass a suspension of fine solid particles or liquid droplets in a gas, such as smoke, oceanic haze, air pollution, and smog. Aerosols may influence climate by either scattering and absorbing radiation, or by acting as condensation nuclei for cloud formation or modifying the properties and lifetime of clouds. the mass of living organic matter (plant and animal) in an ecosystem; biomass also refers to organic matter (living and dead) available on a renewable basis for use as a fuel; biomass includes trees and plants (both terrestrial and aquatic), agricultural crops and wastes, wood and wood wastes, forest and mill residues, animal wastes, livestock operation residues, and some municipal and industrial wastes. asynchronous quanile regression a type of regression used in statistical downscaling. Quantile regression models the relation between a set of predictor variables and specific percentiles (or quantiles) of the response variable. boreal a zone between 50 and 55° and 5 and 70° latitude in the Northern Hemisphere characterized by cool northern temperatures and low rainfall (<20 inches). bagging trees This statistical technique begins with a “regression tree” approach, but recognizes that part of the output error in using a single regression tree comes from the specific selection of an original data set. The bagging trees method uses another statistical technique called “bootstrapping” to create several similar data sets. Regression trees are then produced from these new data sets and results are averaged. carbon dioxide (Co2) ferilizaion barrens Co2-equivalent (CO2-eq) plant communities that occur on sandy soils and that are dominated by grasses, low shrubs, small trees, and scattered large trees. the concentration of carbon dioxide (CO2 ) that would cause the same amount of radiative forcing as a given mixture of CO2 and other forcing components. baselow the condition in which groundwater provides the entire flow of a stream. (During most of the year, streamflow is composed of both groundwater discharge and land surface runoff.) 192 increased plant uptake of CO2 through photosynthesis in response to higher concentrations of atmospheric CO2 . climate normal the arithmetic mean of a climatological element computed over three consecutive decades. convecive storm Convection is a process whereby heat is transported vertically within the atmosphere. Convective storms result from a combination of convection, moisture, and instability. Convective storms can produce thunderstorms, tornadoes, hail, heavy rains, and straight-line winds. GLoSSARY dendriic drainage eastern deciduous forest a stream drainage pattern that resembles the branching pattern of a tree, with tributaries joining larger streams at angles <90°. This type of drainage occurs where the subsurface geology has a uniform resistance to erosion, and therefore little influence on the direction that tributaries take. a forest dominated by trees such as oaks, maples, beech, hickories, and birches that drop their leaves. Evergreen conifers do live in this forest, but are rarely dominant. This forest develops under cold winters (but not as cold as the boreal region to the north), and annual rainfall is higher in this forest than anywhere else in North America except for the subtropical and tropical areas to the south. derecho widespread and long-lived convective windstorm that is associated with a band of rapidly moving showers or thunderstorms characterized by wind gusts that are greater than 57 miles per hour and that may exceed 100 miles per hour (National Oceanic and Atmospheric Administration 2012). ecological processes processes fundamental to the functioning of a healthy and sustainable ecosystem, usually involving the transfer of energy and substances from one medium or trophic level to another. disturbance ecoregion stresses and destructive agents such as invasive species, diseases, and fire; changes in climate and serious weather events such as hurricanes and ice storms; pollution of the air, water, and soil; real estate development of forest lands; and timber harvest. Some of these are caused by humans, in part or entirely; others are not. repetitive pattern of ecosystems associated with commonalities in soil and landform that characterize that larger region. edaphic of or pertaining to soil characteristics. emissions scenario downscaling a method for obtaining high-resolution climate or climate change information from relatively coarseresolution general circulation models (GCMs); involves examining the statistical relationship between past climate data and on-the-ground measurements. a plausible representation of the future development of emissions of greenhouse gases and aerosols that are potentially radiatively active, based on certain demographic, technological, or environmental developments (Intergovernmental Panel on Climate Change [IPCC] 2007). esker driver any natural or human-induced factor that directly or indirectly causes a change in an ecosystem. a serpentine ridge of glacial drift, originally deposited by a meltwater stream running beneath a glacier. dynamical downscaling evapotranspiraion a method for obtaining high-resolution climate or climate change information from relatively coarseresolution general circulation models (GCMs) using a limited-area, high-resolution model (a regional climate model, or RCM) driven by boundary conditions from a GCM to derive smaller-scale information. the sum of evaporation from the soil and transpiration from plants. luvial of, relating to, produced by, or inhabiting a stream or river. 193 GLoSSARY forest type greenhouse efect a classification of forest land based on the dominant species present, as well as associate species commonly occurring with the dominant species. the rise in temperature that the Earth experiences because certain gases in the atmosphere (water vapor, carbon dioxide, nitrous oxide, and methane, for example) absorb and emit energy from the sun. forest-type group based on FIA definitions, a combination of forest types that share closely associated species or site requirements and are generally combined for brevity of reporting. growing season the period in each year when the weather and temperature are right for plants to grow. growing stock fragmentaion a disruption of ecosystem or habitat connectivity, caused by human or natural disturbance, creating a mosaic of successional and developmental stages within or between forested tracts of varying patch size, isolation (distance between patches), and edge length. funcional diversity the value, range, and relative abundance of functional traits in a given ecosystem. a classification of timber inventory that includes live trees of commercial species meeting specified standards of quality or vigor. When associated with volume, this includes only trees ≥5.0 inches in diameter at breast height. habitat those parts of the environment (aquatic, terrestrial, and atmospheric) often typified by a dominant plant form or physical characteristic, on which an organism depends, directly or indirectly, in order to carry out its life processes. fundamental niche the total habitat available to a species based on climate, soils, and land cover type in the absence of competitors, diseases, or predators. general circulaion model (GCM) a mathematical model of the general circulation of a planetary atmosphere or ocean and based on the Navier–Stokes equations on a rotating sphere with thermodynamic terms for various energy sources. glacial drit (ill) unsorted and unstratified drift (typically a heterogeneous mix of sand, silt, clay, gravel, and stones) deposited directly by and underneath a glacier without subsequent reworking by meltwater. hardwood a dicotyledonous tree, usually broad-leaved and deciduous. Hardwoods can be split into soft hardwoods (red maple, paper birch, quaking aspen, and American elm) and hard hardwoods (sugar maple, yellow birch, black walnut, and oaks). impact model simulations of impacts on trees, animals, and ecosystems; these models use GCM projections as inputs, and include additional inputs such as tree species, soil types, and life history traits of individual species. importance value an index of the relative abundance of a species in a given community (0 = least abundant, 50 = most abundant). 194 GLoSSARY industrially owned forest mesophicaion land owned by forest product companies that harvest and market timber. a process “whereby microenvironmental conditions (cool, damp, and shaded conditions; less flammable fuel beds) continually improve for shade-tolerant mesophytic species and deteriorate for shadeintolerant, fire-adapted species” (Nowacki and Abrams 2008: 123). intensity amount of precipitation falling per unit of time. kame a short ridge or mound of stratified drift deposited from a retreating glacier. karst an area of irregular limestone (calcium carbonate) in which erosion has produced fissures, sinkholes, underground streams, and caverns. Most caves are formed below the water table, resulting in stalactites and stalagmites. ketle a depression left in a mass of glacial drift, formed by the melting of an isolated block of glacial ice. model error uncertainty caused by a lack of complete understanding of some climate processes, or by the inability of models to pick up small-scale but influential climate processes. model reliability score for the Tree Atlas: a “tri-model” approach to assess reliability of model predictions for each species, classified as high, medium, or low, depending on the assessment of the stability of the bagged trees and the R2 in RandomForest (Iverson et al. 2008b: 392). modifying factor Kyoto Protocol Adopted at the 1997 Third Session of the Conference of Parties to the UN Framework Convention on Climate Change in Kyoto, Japan, it contains legally binding commitments to reduce anthropogenic greenhouse gas emissions by at least 5 percent below 1990 levels in the period 2008-2012 (IPCC 2007). environmental variables (e.g., site conditions, interspecies competition, disturbance, dispersal ability) that influence the way a tree may respond to climate change. moraine an accumulation of boulders, stones, or other debris carried and deposited by a glacier. lacustrine pertaining to or formed in a lake. nonindustrial private landowners mass wasing an ownership class of private lands where the owner does not operate wood-using plants. movement of water and other materials as controlled by gravity; occurs on slopes under influence of gravitational stress. Gravity pulls on a mass until a critical shear-failure point is reached; thus, the greater the slope, the more mass wasting. mesic northern hardwoods forest type with wet-mesic to dry-mesic soils, medium to high soil nutrient level, and supporting tree species such as sugar maple (dominant), basswood, hemlock, yellow birch, ironwood, red maple, and white ash. pertaining to sites or habitats characterized by intermediate (moist, but not wet or dry) soil moisture conditions. 195 GLoSSARY orographic liting pulpwood the process in which an air mass is forced from a low elevation to a higher elevation. Adiabatic cooling can subsequently raise the relative humidity to 100 percent, resulting in clouds and precipitation. roundwood, whole-tree chips, or wood residues used for the production of wood pulp for making paper and paperboard products. parcelizaion radiaive forcing peak low the change in net irradiance between different layers of the atmosphere. A positive forcing (more incoming energy) tends to warm the system; a negative forcing (more outgoing energy) tends to cool it. Causes include changes in solar radiation or concentrations of radiatively active gases and aerosols. the maximum instantaneous discharge of a stream or river at a given location. RandomForests the subdivision of a single forest ownership into two or more ownerships. Parcelization may result in fragmentation if habitat is altered under new ownership. phenology the timing of natural events such as the date that migrating birds return, the first flower dates for plants, and the date on which a lake freezes in the autumn or opens in the spring. Also refers to the study of this subject. RandomForests is a statistical technique similar to bagging trees in that it also uses bootstrapping to construct multiple regression trees. The difference is that each tree is produced with a random subset of predictors. Typically, 500 to 2,000 trees are produced and the results are aggregated by averaging. This technique eliminates the possibility of overfitting data. process model a model that relies on computer simulations based on mathematical representations of physical and biological processes that interact over space and time. Real Estate Investment Trust (REIT) projecion realized niche a model-derived estimate of future climate, and the pathway leading to it. the portion of potential habitat that a species occupies; usually it is less than what is available because of predation, disease, and competition with other species. proxy a figure or data source that is used as a substitute for another value in a calculation. Ice and sediment cores, tree rings, and pollen fossils are all examples of things that can be analyzed to infer past climate. The size of rings and the isotopic ratios of elements (e.g., oxygen, hydrogen, and carbon) in rings and other substrates allow scientists to infer climate and timing. 19 Considered private, nonindustrial landowners, REITS own and operate large acreages of timberland. recharge the natural process of movement of rainwater from land areas or streams through permeable soils into water-holding rocks that provide underground storage (i.e., aquifers). GLoSSARY refugia scenario locations and habitats that support populations of organisms that are limited to small fragments of their previous geographic range. a coherent, internally consistent, and plausible description of a possible future state of the world. It is not a forecast; rather, each scenario is one alternative image of how the future can unfold. A projection may serve as the raw material for a scenario, but scenarios often require additional information (IPCC 2007). resampling a method to resize or change the resolution of a data grid in geographic information systems. Resampling should not be confused with downscaling. Resampling is performed only on grids that are larger than the original cell size. senescence roundwood the process of aging in plants. Leaf senescence causes leaves of deciduous trees to change color in autumn. logs, bolts, and other round timber generated from harvesting trees for industrial or consumer use. signiicant trend runof that part of the precipitation that appears in surface streams. It is the same as streamflow unaffected by artificial diversions or storage. saw log a log meeting minimum standards of diameter, length, and defect, including logs at least 8 feet long, sound and straight, and with a minimum diameter inside bark of  inches for softwoods and 8 inches for hardwoods, or meeting other combinations of size and defect specified by regional standards. sawimber a live tree of commercial species containing at least a 12-foot saw log or two noncontiguous 8-foot or longer saw logs, and meeting specifications for form; softwoods must be at least 9 inches, and hardwoods must be at least 11 inches, respectively, in diameter outside the bark. significant trends are least-squares regression pvalues of observed climate trends. In this report, significant trends (p < 0.10) are shown by stippling on maps of observed climate trends. Where no stippling appears (p > 0.10), observed trends have a higher probability of being due to chance alone (Girvetz et al. 2009). snowpack layers of accumulated snow that usually melts during warmer months. sotwood a coniferous tree, usually evergreen, having needles or scale-like leaves. species distribuion model a model that uses statistical relationships to project future change. 197 GLoSSARY staisical downscaling a method for obtaining high-resolution climate or climate change information from relatively coarseresolution general circulation models (GCMs) by deriving statistical relationships between observed small-scale (often station level) variables and larger(GCM-) scale variables. Future values of the largescale variables obtained from GCM projections of future climate are then used to drive the statistical relationships and so estimate the smaller-scale details of future climate. Timber Investment Management Organizaion (TIMO) Considered private, nonindustrial landowners, TIMOs act as investment managers for clients who own timberlands as partnership shares. topkill death of aboveground tree stem and branches. transpiraion liquid water phase change occurring inside plants with the vapor diffusing to the atmosphere. stormlow runoff that occurs due to a heavy precipitation event. uncertainty streamlow a term used to describe the range of possible values around a best estimate, sometimes expressed in terms of probability or likelihood. discharge that occurs in a natural surface stream course whether or not it is diverted or regulated. threat a source of danger or harm. 198 vulnerability susceptibility to a threat. LiTERATuRE CiTED Abrams, M.D. 1992. Fire and the development of oak forests. BioScience. 42(5): 34-353. Abrams, M.D. 1998. The red maple paradox. BioScience. 48(5): 355. Abrams, M.D. 2003. Where has all the white oak gone? BioScience. 53(10): 927-939. Abrams, M.D.; Copenheaver, C.A.; Black, B.A.; van de Gevel, S. 2001. Dendroecology and climatic impacts for a relict, old-growth, bog forest in the Ridge and Valley Province of central Pennsylvania, USA. Canadian Journal of Botany. 79(1): 58-9. Abrams, M.D.; Johnson, S.E. 2012. Long-term impacts of deer exclosures on mixed-oak forest composition at the Valley Forge National Historical Park, Pennsylvania, USA. The Journal of the Torrey Botanical Society. 139(2): 17-180. Abrams, M.D.; Nowacki, G.J. 1992. Historical variation in fire, oak recruitment, and postlogging accelerated succession in central Pennsylvania. The Journal of the Torrey Botanical Society. 119(1): 19-28. Abrams, M.D.; Orwig, D.A.; Demeo, T.E. 1995. Dendroecological analysis of successional dynamics for a presettlement-origin whitepine-mixed-oak forest in the southern Appalachians, USA. Journal of Ecology. 83: 123-133. Abrams, M.D.; Ruffner, C.M.; DeMeo, T.E. 1998. Dendroecology and species co-existence in an old-growth Quercus-Acer-Tilia talus slope forest in the central Appalachians, USA. Forest Ecology and Management. 10(1): 9-18. Abrams, M.D.; Scott, M.L. 1989. Disturbancemediated accelerated succession in two Michigan forest types. Forest Science. 35(1): 42-49. Ainsworth, E.A.; Long, S.P. 2005. What have we learned from 15 years of free-air CO2 enrichment (FACE)? A meta-analytic review of the responses of photosynthesis, canopy properties and plant production to rising CO2 . New Phytologist. 15(2): 351-372. Ainsworth, E.A.; Rogers, A. 2007. The response of photosynthesis and stomatal conductance to rising [CO2 ]: mechanisms and environmental interactions. Plant, Cell & Environment. 30(3): 258-270. Aldrich, S.R.; Lafon, C.W.; Grissino-Mayer, H.D.; DeWeese, G.G.; Hoss, J.A. 2010. Three centuries of fire in montane pine-oak stands on a temperate forest landscape. Applied Vegetation Science. 13(1): 3-4. Alpert, P.; Bone, E.; Holzapfel, C. 2000. Invasiveness, invasibility and the role of environmental stress in the spread of nonnative plants. Perspectives in Plant Ecology, Evolution and Systematics. 3(1): 52-. American Tree Farm System [ATFS]. 2014. State forestry fact sheet (Ohio, West Virginia, and Maryland). Available at https://www. treefarmsystem.org/about-maryland-tree-farmprogram. (Accessed January 14, 2014). Anderegg, W.R.; Prall, J.W.; Harold, J.; Schneider, S.H. 2010. Expert credibility in climate change. Proceedings of the National Academy of Sciences. 107(27): 12107-12109. 199 LiTERATuRE CiTED Anderson, M.; Ferree, C. 2010. Conserving the stage: climate change and the geophysical underpinnings of species diversity. PLoS ONE. 5(7): e11554. Anderson, M.G.; Clark, M.; Sheldon, A.O. 2012. Resilient sites for terrestrial conservation in the Northeast and Mid-Atlantic Region. Boston, MA: The Nature Conservancy, Eastern Conservation Science. 18 p. Anderson, R.L. 1995. How to identify and control Hypoxylon cankers of oaks and other hardwoods. Protection Report R8-PR 29. Atlanta, GA: U.S. Department of Agriculture, Forest Service, Southern Region. 9 p. Available at http://www.fs.fed.us/r8/foresthealth/publications/ r8_pr_29.pdf. (Accessed August 8, 2014). Andreadis, K.M.; Lettenmaier, D.P. 200. Trends in 20th century drought over the continental United States. Geophysical Research Letters. 33(10): L10403. Andrew, N.R.; Hill, S.J.; Binns, M.; Bahar, M.H.; Ridley, E.V.; Jung, M.-P.; Fyfe, C.; Yates, M.; Khusro, M. 2013. Assessing insect responses to climate change: What are we testing for? Where should we be heading? PeerJ. 1: e11. Appalachian Mountains Joint Venture Management Board. 2008. Implementation plan for the Appalachian Mountains Joint Venture: a foundation for all-bird conservation in the region. Smith, B., ed. Frankfort, KY: Appalachian Mountains Joint Venture. 92 p. Available at http:// amjv.org/documents/implementation_plan.pdf. (Accessed August 21, 2014). Ashfaq, M.; Bowling, L.C.; Cherkauer, K.; Pal, J.S.; Diffenbaugh, N.S. 2010. Influence of climate model biases and daily-scale temperature and precipitation events on hydrological impacts assessment: a case study of the United States. Journal of Geophysical Research. 115(D14): D1411. 200 Babin-Fenske, J.; Anand, M. 2011. Agent-based simulation of effects of stress on forest tent caterpillar (Malacosoma disstria Hübner) population dynamics. Ecological Modelling. 222(14): 251-259. Badeck, F.W.; Bondeau, A.; Böttcher, K.; Doktor, D.; Lucht, W.; Schaber, J.; Sitch, S. 2004. Responses of spring phenology to climate change. New Phytologist. 12(2): 295-309. Bader, D.C.; Covey, C.; Gutkowski, W.J., Jr.; Held, I.M.; Kunkel, K.E.; Miller, R.L.; Tokmakian, R.T.; Zhang, M.H. 2008. Climate models: an assessment of strengths and limitations. U.S. Climate Change Science Program Synthesis and Assessment Product 3.1. Washington, DC: U.S. Department of Energy, Office of Biological and Environmental Research. 124 p. Available at http://pubs.giss.nasa.gov/abs/ba09200l.html. (Accessed August 20, 2014). Bar Massada, A.; Radeloff, V.C.; Stewart, S.I.; Hawbaker, T.J. 2009. Wildfire risk in the wildland-urban interface: a simulation study in northwestern Wisconsin. Forest Ecology and Management. 258(9): 1990-1999. Bartomeus, I.; Ascher, J.S.; Wagner, D.; Danforth, B.N.; Colla, S.; Kornbluth, S.; Winfree, R. 2011. Climate-associated phenological advances in bee pollinators and bee-pollinated plants. Proceedings of the National Academy of Sciences. 108(51): 2045-2049. Beniston, M. 2003. Climatic change in mountain regions: a review of possible impacts. Climatic Change. 59: 5-31. Benson, B.J.; Magnuson, J.J.; Jensen, O.P.; Card, V.M.; Hodgkins, G.; Korhonen, J.; Livingstone, D.M.; Stewart, K.M.; Weyhenmeyer, G.A.; Granin, N.G. 2012. Extreme events, trends, and variability in Northern Hemisphere lakeice phenology (1855-2005). Climatic Change. 112(2): 299-323. LiTERATuRE CiTED Birch, T.W.; Wharton, E.H. 1982. Land use change in Ohio, 1952 to 1979. Resour. Bull. NE-70. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northeastern Forest Experiment Station. Available at http://www. fs.fed.us/ne/newtown_square/publications/ resource_bulletins/pdfs/scanned/OCR/ne_rb70. pdf. (Accessed August 21, 2014). Birdsey, R.; Pregitzer, K.; Lucier, A. 200. Forest carbon management in the United States: 1600-2100. American Society of Agronomy: 35(4): 141-149. Black, B.A.; Abrams, M.D. 2001. Influences of Native Americans and surveyor biases on metes and bounds witness-tree distribution. Ecology. 82(9): 2574-258. Blunden, J.; Arndt, D.S., eds. 2012. State of the climate in 2011. Bulletin of the American Meteorological Society. 93(7): S1-S24. Boer, G. 2009. Changes in interannual variability and decadal potential predictability under global warming. Journal of Climate. 22(11): 3098-3109. Boerner, R.E.J. 200. Unraveling the Gordian Knot: interactions among vegetation, topography, and soil properties in the central and southern Appalachians. The Journal of the Torrey Botanical Society. 133(2): 321-31. Bonan, G.B. 2008. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science. 320(5882): 1444-1449. Bonnin, G.M.; Maitaria, K.; Yekta, M. 2011. Trends in rainfall exceedances in the observed record in selected areas of the United States. Journal of the American Water Resources Association. 47(): 1173-1182. Boucher, D.; Rodick, C.; Bailey, J.; Snitzer, J.; Kyde, K.; Prudden, B. 2005. Hurricane Isabel and the forests of the mid-Atlantic Piedmont and Blue Ridge: short-term impacts and long-term implications. In: Sellner, K. G., ed. Hurricane Isabel in perspective. Edgewater, MD: Chesapeake Research Consortium. 05-10. Available at http://www.chesapeake.org/pubs/ Isabel/Boucher%20et%20al.pdf. (Accessed August 20, 2014). Box, G.E.P.; Draper, N.R. 1987. Empirical modelbuilding and response surfaces. Wiley Series in probability and mathematical statistics. New York, NY: John Wiley & Sons. 9 p. Bradley, B.A.; Oppenheimer, M.; Wilcove, D.S. 2009. Climate change and plant invasions: restoration opportunities ahead? Global Change Biology. 15(): 1511-1521. Bradley, B.A.; Wilcove, D.S.; Oppenheimer, M. 2010. Climate change increases risk of plant invasion in the eastern United States. Biological Invasions. 12(): 1855-1872. Brooks, M.; Lusk, M. 2008. Fire management and invasive plants: a handbook. Arlington, VA: US Fish and Wildlife Service. 27 p. Available at http://www.fws.gov/invasives/pdfs/usfws_ firemgtandinvasivesplants_a_handbook.pdf. (Accessed August 20, 2014). Brose, P.; Van Lear, D. 1999. Effects of seasonal prescribed fires on residual overstory trees in oak-dominated shelterwood stands. Southern Journal of Applied Forestry. 23(2): 88-93. Brose, P.H.; Dey, D.C.; Phillips, R.J.; Waldrop, T.A. 2013. A meta-analysis of the fire-oak hypothesis: Does prescribed burning promote oak reproduction in eastern North America? Forest Science. 59(3): 322-334. Borken, W.; Matzner, E. 2009. Reappraisal of drying and wetting effects on C and N mineralization and fluxes in soils. Global Change Biology. 15(4): 808-824. 201 LiTERATuRE CiTED Brose, P.H.; Gottschalk, K.W.; Horsley, S.B.; Knopp, P.D.; Kochenderfer, J.N.; McGuinness, B.J.; Miller, G.W.; Ristau, T.E.; Stoleson, S.H.; Stout, S.L. 2008. Prescribing regeneration treatments for mixed-oak forests in the Mid-Atlantic region. Gen. Tech. Rep. NRS-33. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 108 p. Brose, P.H.; Waldrop, T.A. 200. Fire and the origin of Table Mountain pine pitch pine communities in the southern Appalachian mountains, USA. Canadian Journal of Forest Research. 3(3): 710-718. Brown, R.L.; Peet, R.K. 2003. Diversity and invasibility of southern Appalachian plant communities. Ecology. 84(1): 32-39. Bruce, O.C.; Smith, A.M. 1921. Soil survey of Allegany County, Maryland. Field Operations of the Bureau of Soils. 103-1090. Available at http://soils.usda.gov/survey/online_surveys/ maryland/alleganyMD1921/alleganyMD1921. pdf. (Accessed August 21, 2014). Brzostek, E.R.; Blair, J.M.; Dukes, J.S.; Frey, S.D.; Hobbie, S.E.; Melillo, J.M.; Mitchell, R.J.; Pendall, E.; Reich, P.B.; Shaver, G.R. 2012. The effect of experimental warming and precipitation change on proteolytic enzyme activity: positive feedbacks to nitrogen availability are not universal. Global Change Biology. 18(8): 217-225. Buenger, B.A. 2003. The impact of wildland and prescribed fire on archaeological resources. Lawrence, KS: University of Kansas. 351 p. Ph.D. dissertation. Available at http:// www.blm.gov/pgdata/etc/medialib/blm/wo/ Planning_and_Renewable_Resources/coop_ agencies/cr_publications.Par.30817.File.dat/ DissertationBuenger_for_merge.pdf. (Accessed August 21, 2014). 202 Burnett, A.W.; Kirby, M.E.; Mullins, H.T.; Patterson, W.P. 2003. Increasing great lake-effect snowfall during the twentieth century: A regional response to global warming? Journal of Climate. 1(21): 3535-3542. Burns, N.M.; Rockwell, D.C.; Bertram, P.E.; Dolan, D.M.; Ciborowski, J.J. 2005. Trends in temperature, secchi depth, and dissolved oxygen depletion rates in the central basin of Lake Erie, 1983-2002. Journal of Great Lakes Research. 31: 35-49. Burns, R.M.; Honkala, B.H. 1990. Silvics of North America: 1. Conifers; 2. Hardwoods. Washington, DC: U.S. Department of Agriculture, Forest Service: 877 p. Bussler, B.H.; Byrnes, W.R.; Pope, P.E.; Chaney, W.R. 1984. Properties of minesoil reclaimed for forest land use. Soil Science Society of America Journal. 48(1): 178-184. Butler, B.J. 2008. Family forest owners of the United States, 2006. Gen. Tech. Rep. NRS-27. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 73 p. Byers, E.; Norris, S. 2011. Climate change vulnerability assessment of species of concern in West Virginia. Project Report. Elkins, WV: West Virginia Division of Natural Resources. 9 p. Available at http://wvdnr.gov/publications/ PDFFiles/ClimateChangeVulnerability.pdf. (Accessed August 20, 2014). Byers, E.A.; Vanderhorst, J.P.; Streets, B.P. 2007. Classification and conservation assessment of high elevation wetland communities in the Allegheny mountains of West Virginia. Elkins, WV: West Virginia Division of Natural Resources, Wildlife Resources Section. 92 p. LiTERATuRE CiTED Campbell, J.L.; Mitchell, M.J.; Groffman, P.M.; Christenson, L.M.; Hardy, J.P. 2005. Winter in northeastern North America: a critical period for ecological processes. Frontiers in Ecology and the Environment. 3(): 314-322. Campbell, J.L.; Ollinger, S.V.; Flerchinger, G.N.; Wicklein, H.; Hayhoe, K.; Bailey, A.S. 2010. Past and projected future changes in snowpack and soil frost at the Hubbard Brook Experimental Forest, New Hampshire, USA. Hydrological Processes. 24(17): 245-2480. Certini, G. 2005. Effects of fire on properties of forest soils: a review. Oecologia. 143(1): 1-10. Chamberlain, J.; Winn, M.; Hammett, A.L. 2009. Connecting non-timber forest products stakeholders to information and knowledge: a case study of an intranet Web site. In: Ashton, S.F.; Hubbard, W.G.; Rauscher, H.M., eds. A Southern Region conference on technology transfer and extension. Gen. Tech. Rep. SRS-11. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station: 87-94. Available at http://www.srs.fs.usda.gov/pubs/gtr/ gtr_srs11/gtr_srs11-04.pdf. (Accessed August 20, 2014). Changnon, D.; Changnon, S.A. 2007. Snowstorm dimensions across the central and eastern United States. Physical Geography. 28(3): 218-232. Changnon, S. 2011a. Temporal distribution of weather catastrophes in the USA. Climatic Change. 10(2): 129-140. Changnon, S.A. 2003. Geographical and temporal variations in thunderstorms in the contiguous United States during the 20th century. Physical Geography. 30(2): 138-152. Changnon, S.A. 2011b. Major damaging convective storms in the United States. Physical Geography. 32(3): 28-294. Chen, I.-C.; Hill, J.K.; Ohlemüller, R.; Roy, D.B.; Thomas, C.D. 2011. Rapid range shifts of species associated with high levels of climate warming. Science. 333(045): 1024-102. Choat, B.; Jansen, S.; Brodribb, T.J.; Cochard, H.; Delzon, S.; Bhaskar, R.; Bucci, S.J.; Feild, T.S.; Gleason, S.M.; Hacke, U.G. 2012. Global convergence in the vulnerability of forests to drought. Nature. 491: 752-75. Chong, S.-K.; Cowsert, P. 1997. Infiltration in reclaimed mined land ameliorated with deep tillage treatments. Soil and Tillage Research. 44(3): 255-24. Christidis, N.; Stott, P.A.; Brown, S.; Karoly, D.J.; Caesar, J. 2007. Human contribution to the lengthening of the growing season during 1950-99. Journal of Climate. 20(21): 5441-5454. Clark, J.S.; Fastie, C.; Hurtt, G.; Jackson, S.T.; Johnson, C.; King, G.A.; Lewis, M.; Jason, L.; Pacala, S.; Prentice, C.; Schupp, E.W.; Webb, T., III; Wyckoff, P. 1998. Reid’s paradox of rapid plant migration. BioScience. 48(1): 13-24. Clatterbuck, W.K.; Kauffman, B.W. 200. Managing oak decline. SP75. Knoxville, TN: University of Tennessee, UT Extension.  p. Available at http://www.uky.edu/Ag/Forestry/ extension/pub/pdf/for99.pdf. (Accessed August 21, 2014). Cleland, D.T.; Freeouf, J.; Keys, J., Jr.; Nowacki, G.J.; Carpenter, C.; McNab, W.H. 2007a. Ecological subregions: sections and subsections of the conterminous United States [1:3,500,000] [CD-ROM]. Online database. Sloan, A.M., cartog. Gen. Tech. Rep. WO-7. Washington, DC: U.S. Department of Agriculture, Forest Service. Available at http://data.fs.usda.gov/geodata/edw/datasets. php?dsetCategory=biota. (Accessed October 22, 2013). 203 LiTERATuRE CiTED Cleland, E.E.; Chuine, I.; Menzel, A.; Mooney, H.A.; Schwartz, M.D. 2007b. Shifting plant phenology in response to global change. Trends in Ecology & Evolution. 22(7): 357-35. ClimateWizard. 2013. Climate Wizard Custom Analysis application. The Nature Conservancy, University of Washington, University of Southern Mississippi, Climate Central, Santa Clara University. Available at http:// climatewizardcustom.org/. (Accessed May 2013). Cole, C.T.; Anderson, J.; Lindroth, R.L.; Waller, D.M. 2010. Rising concentrations of atmospheric CO2 have increased growth in natural stands of quaking aspen (Populus tremuloides). Global Change Biology. 1(8): 218-2197. Collins, R.J.; Carson, W.P. 2003. The fire and oak hypothesis: incorporating the influence of deer browsing and canopy gaps. In: Van Sambeek, J.W.; Dawson, J.O.; Ponder, F., Jr.; Loewenstein, E.F.; Fralish, J.S., eds. Proceedings, 13th Central Hardwoods Forest conference. Gen. Tech. Rep. NC-234. St. Paul, MN: U.S. Department of Agriculture, Forest Service, North Central Research Station: 44-3. Comer, P.; Faber-Langendoen, D.; Evans, R.; Gawler, S.; Josse, C.; Kittel, G.; Menard, S.; Pyne, M.; Reid, M.; Schultz, K.; Snow, K.; Teague, J. 2003. Ecological systems of the United States: a working classification of U.S. terrestrial systems. Arlington, VA: NatureServe. 83 p. Coniglio, M.C.; Stensrud, D.J. 2004. Interpreting the climatology of derechos. Weather and Forecasting. 19(3): 595-05. Cook, J.; Nuccitelli, D.; Green, S.A.; Richardson, M.; Winkler, B.; Painting, R.; Way, R.; Jacobs, P.; Skuce, A. 2013. Quantifying the consensus on anthropogenic global warming in the scientific literature. Environmental Research Letters. 8(024024): 7. 204 Cornett, M.W.; Reich, P.B.; Puettmann, K.J.; Frelich, L.E. 2000. Seedbed and moisture availability determine safe sites for early Thuja occidentalis (Cupressaceae) regeneration. American Journal of Botany. 87(12): 1807-1814. Côté, S.D.; Rooney, T.P.; Tremblay, J.P.; Dussault, C.; Waller, D.M. 2004. Ecological impacts of deer overabundance. Annual Review of Ecology Evolution and Systematics. 35: 113-147. Coumou, D.; Rahmstorf, S. 2012. A decade of weather extremes. Nature Climate Change. 2: 491-49. Dai, A.; Trenberth, K.E.; Qian, T. 2004. A global dataset of Palmer drought severity index for 1870-2002: relationship with soil moisture and effects of surface warming. Journal of Hydrometeorology. 5: 1117-1129. Dale, V.H.; Joyce, L.A.; McNulty, S.; Neilson, R.P.; Ayres, M.P.; Flannigan, M.D.; Hanson, P.J.; Irland, L.C.; Lugo, A.E.; Peterson, C.J.; Simberloff, D.; Swanson, F.J.; Stocks, B.J.; Wotton, B.M. 2001. Climate change and forest disturbances. BioScience. 51(9): 723-734. Daly, C.; Conklin, D.R.; Unsworth, M.H. 2010. Local atmospheric decoupling in complex topography alters climate change impacts. International Journal of Climatology. 30(12): 1857-184. Daniels, A.E.; Morrison, J.F.; Joyce, L.A.; Crookston, N.L.; Chen, S.C.; McNulty, S.G. 2012. Climate projections FAQ. Gen. Tech. Rep. RMRS-GTR-277WWW. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 32 p. Davis, M.B. 1983. Quaternary history of deciduous forests of eastern North-America and Europe. Annals of the Missouri Botanical Garden. 70(3): 550-53. LiTERATuRE CiTED Davis, M.B.; Shaw, R.G. 2001. Range shifts and adaptive responses to quaternary climate change. Science. 292(5517): 73-79. Davis, M.B.; Shaw, R.G.; Etterson, J.R. 2005. Evolutionary responses to changing climate. Ecology. 8(7): 1704-1714. DeGaetano, A.T.; Allen, R.J. 2002. Trends in twentieth-century temperature extremes across the United States. Journal of Climate. 15(22): 3188-3205. Delpla, I.; Jung, A.-V.; Baures, E.; Clement, M.; Thomas, O. 2009. Impacts of climate change on surface water quality in relation to drinking water production. Environment International. 35(8): 1225-1233. Delworth, T.L.; Broccoli, A.J.; Rosati, A.; Stouffer, R.J.; Balaji, V.; Beesley, J.A.; Cooke, W.F.; Dixon, K.W.; Dunne, J.; Dunne, K.A.; Durachta, J.W.; Findell, K.L.; Ginoux, P.; Gnanadesikan, A.; Gordon, C.T.; Griffies, S.M.; Gudgel, R.; Harrison, M.J.; Held, I.M.; Hemler, R.S.; Horowitz, L.W.; Klein, S.A.; Knutson, T.R.; Kushner, P.J.; Langenhorst, A.R.; Lee, H.C.; Lin, S.J.; Lu, J.; Malyshev, S.L.; Milly, P.C.D.; Ramaswamy, V.; Russell, J.; Schwarzkopf, M.D.; Shevliakova, E.; Sirutis, J.J.; Spelman, M.J.; Stern, W.F.; Winton, M.; Wittenberg, A.T.; Wyman, B.; Zeng, F.; Zhang, R. 200. GFDL’s CM2 global coupled climate models. Part I: formulation and simulation characteristics. Journal of Climate. 19(5): 43-74. DeSantis, R.D.; Moser, W.K.; Huggett, R.J.; Li, R.; Wear, D.N.; Miles, P.D. 2013. Modeling the effects of emerald ash borer on forest composition in the Midwest and Northeast United States. Gen. Tech. Rep. NRS-112. Newtown Square, PA: U.S. Department of Agriculture, Northern Research Station. 23 p. Diffenbaugh, N.S.; Pal, J.S.; Trapp, R.J.; Giorgi, F. 2005. Fine-scale processes regulate the response of extreme events to global climate change. Proceedings of the National Academy of Sciences of the United States of America. 102(44): 15774-15778. Diffenbaugh, N.S.; Trapp, R.J.; Brooks, H. 2008. Does global warming influence tornado activity? Eos, Transactions American Geophysical Union. 89(53): 553-554. Dijak, W. 2013. Landscape Builder: software for the creation of initial landscapes for LANDIS from FIA data. Computational Ecology and Software. 3(2): 17-25. Dobrowski, S.Z. 2011. A climatic basis for microrefugia: the influence of terrain on climate. Global Change Biology. 17(2): 1022-1035. Doran, P.T.; Zimmerman, M.K. 2009. Examining the scientific consensus on climate change. Eos, Transactions American Geophysical Union. 90(3): 22. Dragoni, D.; Rahman, A.F. 2012. Trends in fall phenology across the deciduous forests of the Eastern USA. Agricultural and Forest Meteorology. 157(15): 9-105. Drake, B.G.; Gonzalez-Meler, M.A.; Long, S.P. 1997. More efficient plants: A consequence of rising atmospheric CO2 ? Annual Review of Plant Physiology and Plant Molecular Biology. 48(1): 09-39. Driscoll, C.T.; Han, Y.-J.; Chen, C.Y.; Evers, D.C.; Lambert, K.F.; Holsen, T.M.; Kamman, N.C.; Munson, R.K. 2007. Mercury contamination in forest and freshwater ecosystems in the northeastern United States. BioScience. 57(1): 17-28. Diffenbaugh, N.S.; Ashfaq, M. 2010. Intensification of hot extremes in the United States. Geophysical Research Letters. 37(15): L15701. 205 LiTERATuRE CiTED Drohan, P.; Brittingham, M.; Bishop, J.; Yoder, K. 2012a. Early trends in landcover change and forest fragmentation due to shale-gas development in Pennsylvania: a potential outcome for the north central Appalachians. Environmental Management. 49: 101-1075. Drohan, P.J.; Finley, J.C.; Roth, P.; Schuler, T.M.; Stout, S.L.; Brittingham, M.C.; Johnson, N.C. 2012b. Oil and gas impacts on forest ecosystems: findings gleaned from the 2012 Goddard Forum at Penn State University. Environmental Practice. doi:10.10170S140412000300. Dukes, J.S.; Pontius, J.; Orwig, D.; Garnas, J.R.; Rodgers, V.L.; Brazee, N.; Cooke, B.; Theoharides, K.A.; Stange, E.E.; Harrington, R.; Ehrenfeld, J.; Gurevitch, J.; Lerdau, M.; Stinson, K.; Wick, R.; Ayres, M. 2009. Responses of insect pests, pathogens, and invasive plant species to climate change in the forests of northeastern North America: What can we predict? Canadian Journal of Forest Research. 39(2): 231-248. Dunnington, G. 2011. The potential impacts of climate change on recreation in Vermont. Climate Change Adaptation White Paper Series. Montpelier, VT: Vermont Agency of Natural Resources. 5 p. Available at http:// www.anr.state.vt.us/anr/climatechange/Pubs/ VTCCAdaptRecreation.pdf. (Accessed August 20, 2014). Dye, C.R. 2013. Report on managed timberland program. 4 p. Charleston, WV: West Virginia Legislature. Available at http://www.legis.state. wv.us/legisdocs/reports/agency/F03_FY_2013_ 2259.pdf. (Accessed August 21, 2014). Dyer, J.M. 2001. Using witness trees to assess forest change in southeastern Ohio. Canadian Journal of Forest Research. 31(10): 1708-1718. 20 Easterling, D.R. 2002. Recent changes in frost days and the frost-free season in the United States. Bulletin of the American Meteorological Society. 83(9): 1327-1332. Ebi, K.L.; Meehl, G.A. 2007. The heat is on: climate change & heatwaves in the Midwest. In: Ebi, K.L.; Meel, G.A., eds. Regional impacts of climate change: four case studies in the United States. Arlington, VA: Pew Center on Global Climate Change: 8-21. Available at http://www. c2es.org/docUploads/Regional-Impacts-Midwest. pdf. (Accessed August 21, 2014). Ehrenfeld, J.G. 2010. Ecosystem consequences of biological invasions. Annual Review of Ecology, Evolution, and Systematics. 41: 59-80. Eisenbies, M.; Aust, W.; Burger, J.; Adams, M.B. 2007. Forest operations, extreme flooding events, and considerations for hydrologic modeling in the Appalachians - a review. Forest Ecology and Management. 242(2): 77-98. Elliott, K.J.; Vose, J.M.; Jackson, W.A. 2013. Effects of future sulfate and nitrate deposition scenarios on Linville Gorge and Shining Rock Wildernesses. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 30 p. Available at http://www.srs.fs.fed. us/pubs/gtr/gtr_srs181.pdf. (Accessed August 21, 2014). Emery, M.R.; Barron, E.S. 2010. Using local ecological knowledge to assess morel decline in the U.S. Mid-Atlantic region. Economic Botany. 4(3): 205-21. Felzer, B.S.; Reilly, J.M.; Melillo, J.M.; Kicklighter, D.W.; Wang, C.; Prinn, R.G.; Sarofim, M.C.; Zhuang, Q. 2003. Past and future effects of ozone on net primary production and carbon sequestration using a global biogeochemical model. Report No. 103. Cambridge, MA: MIT Joint Program on the Science and Policy of Global Change. Available at http://dspace.mit. edu/bitstream/handle/1721.1/4053/MITJPSPGC_ Rpt_103.pdf?sequence=2. (Accessed August 21, 2014). LiTERATuRE CiTED Feng, S.; Hu, Q. 2007. Changes in winter snowfall/ precipitation ratio in the contiguous United States. Journal of Geophysical Research. 112(D15): D15109. Fernholz, K.; Bowyer, J.; Howe, J. 2007. TIMOs and REITs: what, why, and how they might impact sustainable forestry. Minneapolis, MN: Dovetail Partners, Inc. 14 p. Available at http:// www.iatp.org/files/181_2_98730.pdf. (Accessed August 21, 2014). Finzi, A.C.; Sinsabaugh, R.L.; Long, T.M.; Osgood, M.P. 200. Microbial community responses to atmospheric carbon dioxide enrichment in a warm-temperate forest. Ecosystems. 9: 215-22. Fischer, A.P.; Paveglio, T.; Carroll, M.; Murphy, D.; Brenkert-Smith, H. 2013. Assessing social vulnerability to climate change in human communities near public forests and grasslands: a framework for resource managers and planners. Journal of Forestry. 111(5): 357-35. Fitzjarrald, D.R.; Acevedo, O.C.; Moore, K.E. 2001. Climatic consequences of leaf presence in the eastern United States. Journal of Climate. 14(4): 598-14. Forest2Market. 2013. SFI certification by region and state. Available at http://www.forest2market. com/uploads/Legacy/Certification_by_Region_ and_State.pdf. (Accessed August 21, 2014). Foster, J.R.; Burton, J.I.; Forrester, J.A.; Liu, F.; Muss, J.D.; Sabatini, F.M.; Scheller, R.M.; Mladenoff, D.J. 2010. Evidence for a recent increase in forest growth is questionable. Proceedings of the National Academy of Sciences. 107(21): E8-87. Franklin, J.F.; Mitchell, R.J.; Palik, B. 2007. Natural disturbance and stand development principles for ecological forestry. Gen. Tech. Rep. NRS19. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 48 p. Fraser, K.C.; Silverio, C.; Kramer, P.; Mickle, N.; Aeppli, R.; Stutchbury, B.J.M. 2013. A trans-hemispheric migratory songbird does not advance spring schedules or increase migration rate in response to record-setting temperatures at breeding sites. PLoS ONE. 8(5): e4587. Frelich, L.E.; Peterson, R.O.; Dovciak, M.; Reich, P.B.; Vucetich, J.A.; Eisenhauer, N. 2012. Trophic cascades, invasive species and bodysize hierarchies interactively modulate climate change responses of ecotonal temperate-boreal forest. Philosophical Transactions of the Royal Society B: Biological Sciences. 37(105): 2955-291. Fridley, J.D. 2009. Downscaling climate over complex terrain: high finescale (< 1000 m) spatial variation of near-ground temperatures in a Montane forested landscape (Great Smoky Mountains). Journal of Applied Meteorology and Climatology. 48(5): 1033-1049. Friedland, A.J.; Gregory, R.A.; Karenlampi, L.; Johnson, A.H. 1984. Winter damage to foliage as a factor in red spruce decline. Canadian Journal of Forest Research. 14(): 93-95. Froese, R.; Hyslop, M.; Miller, C.; Garmon, B.; McDiarmid, H., Jr.; Shaw, A.; Leefers, L.A.; Lorenzo, M.; Brown, S.; Shy, M. 2007. Largetract forestland ownership change: land use, conservation, and prosperity in Michigan’s Upper Peninsula. Ann Arbor, MI: National Wildlife Federation. 54 p. Available at http:// forestlands.mtu.edu/products/UP_Corporate_ Forest_Land_Report.pdf. (Accessed August 21, 2014). Fryer, J.L. 2011. Microstegium vimineum. Fire effects information system. [Online]. Missoula, MT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory. Available at http://www. fs.fed.us/database/feis/plants/graminoid/micvim/ all.html. (Accessed August 21, 2014). 207 LiTERATuRE CiTED Furedi, M.; Leppo, B.; Kowalski, M.; Davis, T.; Eichelberger, B. 2011. Identifying species in Pennsylvania potentially vulnerable to climate change. Pittsburgh, PA: Pennsylvania Natural Heritage Program, Western Pennsylvania Conservancy. Available at http://www. naturalheritage.state.pa.us/ccvi/ccvi_final_report. pdf. (Accessed August 21, 2014). Gómez-Aparicio, L.; Pérez Ramos, I.M.; Mendoza, I.; Matías, L.; Quero, J.; Castro, J.; Zamora, R.; Marañón, T. 2008. Oak seedling survival and growth along resource gradients in Mediterranean forests: implications for regeneration in current and future environmental scenarios. Oikos. 117(11): 183-199. Gange, A.; Gange, E.; Sparks, T.; Boddy, L. 2007. Rapid and recent changes in fungal fruiting patterns. Science. 31(5821): 71-71. Grafton, E. 2013. Dirty dozen. West Virginia Division of Natural Resources, Wildlife Resources Section. Available at http://www. wvdnr.gov/Wildlife/DirtyDozen.shtm. (Accessed August 21, 2014). Garroway, C.J.; Bowman, J.; Holloway, G.L.; Malcolm, J.R.; Wilson, P.J. 2011. The genetic signature of rapid range expansion by flying squirrels in response to contemporary climate warming. Global Change Biology. 17(5): 170-179. Georgakakos, A.; Fleming, P.; Dettinger, M.; Peters-Lidard, C.; Richmond, T.; Reckhow, K.; White, K.; Yates, D. 2014. Chapter 3: water resources. In: Melillo, J.M.; Richmond, T.; Yohe, G., eds. Climate change impacts in the United States: the third National Climate Assessment. Washington, DC: U.S. Global Change Research Program: 9-112. Gibson, W.P.; Daly, C.; Kittel, T.; Nychka, D.; Johns, C.; Rosenbloom, N.; McNab, A.; Taylor, G. 2002. Development of a 103-year high-resolution climate data set for the conterminous United States. Proceedings of the 13th American Meteorological Society Conference on Applied Climatology: 181-183. Girvetz, E.H.; Zganjar, C.; Raber, G.T.; Maurer, E.P.; Kareiva, P.; Lawler, J.J. 2009. Applied climatechange analysis: the Climate Wizard tool. PLoS ONE. 4(12): e8320. Glick, P.; Stein, B.A.; Edelson, N.A. 2011. Scanning the conservation horizon: a guide to climate change vulnerability assessment. Washington, DC: National Wildlife Federation. 17 p. 208 Gray, D.M.; Dighton, J. 200. Mineralization of forest litter nutrients by heat and combustion. Soil Biology and Biochemistry. 38(): 149-1477. Gray, D.R. 2008. The relationship between climate and outbreak characteristics of the spruce budworm in eastern Canada. Climatic Change. 87(3-4): 31-383. Groffman, P.M.; Kareiva, P.; Carter, S.; Grimm, N.B.; Lawler, J.; Mack, M.; Matzek, V.; Tallis, H. 2014. Chapter 8: ecosystems, biodiversity, and ecosystem services. In: Melillo, J.M.; Richmond, T.; Yohe, G., eds. Climate change impacts in the United States: the third National Climate Assessment. Washington, DC: U.S. Global Change Research Program: 195-219. Groisman, P.Y.; Knight, R.W.; Easterling, D.R.; Karl, T.R.; Hegerl, G.C.; Razuvaev, V.N. 2005. Trends in intense precipitation in the climate record. Journal of Climate. 18(9): 132-1350. Groisman, P.Y.; Knight, R.W.; Karl, T.R. 2012. Changes in intense precipitation over the central United States. Journal of Hydrometeorology. 13(1): 47-. LiTERATuRE CiTED Groisman, P.Y.; Knight, R.W.; Karl, T.R.; Easterling, D.R.; Sun, B.; Lawrimore, J.H. 2004. Contemporary changes of the hydrological cycle over the contiguous United States: trends derived from in situ observations. Journal of Hydrometeorology. 5(1): 4-85. Gutowski, W.J.; Hegerl, G.C.; Holland, G.J.; Knutson, T.R.; Mearns, L.O.; Stouffer, R.J.; Webster, P.J.; Wehner, M.F.; Zwiers, F.W. 2008. Causes of observed changes in extremes and projections of future changes. In: Karl, T.R.; Meehl, G.A.; Miller, C.D.; Hassol, S.J.; Waple, A.M.; and Murray, W.L., eds. Weather and climate extremes in a changing climate. Regions of focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. Washington, DC: U.S. Climate Change Science Program and the Subcommittee on Global Change Research: 81-11. Hansen, J.; Sato, M.; Ruedy, R. 2012. Perception of climate change. Proceedings of the National Academy of Sciences. 109(37): E2415-E2423. Hanson, P.J.; Weltzin, J.F. 2000. Drought disturbance from climate change: response of United States forests. The Science of the Total Environment. 22(3): 205-220. Hartley, S. 1999. Winter Atlantic climate and snowfall in the south and central Appalachians. Physical Geography. 20(1): 1-13. Hayhoe, K. 2010a. 1/8 degree CONUS daily downscaled climate projections *PROVISIONAL*. USGS Geodata Portal [Online]: Texas Tech University, Center for Integrated Data Analytics. Available at http://cida. usgs.gov/thredds/catalog.html?dataset=cida.usgs. gov/thredds/dcp/conus. (Accessed July 11, 2011). Hayhoe, K.; Stoner, A.; Yang, X.; Crow, C.; Swaminathan, R.; Scott-Fleming, I.; Ryu, J-H.; Gelca, R.; Swain, S. 2013. Development and dissemination of a highresolution national climate change dataset. Texas Tech University High Performance Computing Center. 497 p. Available at http://cida. usgs.gov/thredds/fileServer/dcp/files/Hayhoe_ USGS_downscaled_database_final_report.pdf. (Accessed August 21, 2014). Hayhoe, K.; Wake, C.P.; Huntington, T.G.; Luo, L.F.; Schwartz, M.D.; Sheffield, J.; Wood, E.; Anderson, B.; Bradbury, J.; DeGaetano, A.; Troy, T.J.; Wolfe, D. 2007. Past and future changes in climate and hydrological indicators in the US Northeast. Climate Dynamics. 28(4): 381-407. Hayhoe, K.A. 2010b. A standardized framework for evaluating the skill of regional climate downscaling techniques. Urbana, IL: University of Illinois at Urbana-Champaign. 153 p. Ph.D. dissertation. Available at http://www.snap.uaf. edu/attachments/1_Hayhoe_Katharine.pdf. (Accessed August 21, 2014). He, H.S.; Hao, Z.; Mladenoff, D.J.; Shao, G.; Hu, Y.; Chang, Y. 2005. Simulating forest ecosystem response to climate warming incorporating spatial effects in northeastern China. Journal of Biogeography. 32(12): 2043-205. He, H.S.; Mladenoff, D.J. 1999. The effects of seed dispersal on the simulation of long-term forest landscape change. Ecosystems. 2(4): 308-319. He, H.S.; Mladenoff, D.J.; Crow, T.R. 1999. Linking an ecosystem model and a landscape model to study forest species response to climate warming. Ecological Modelling. 114(2): 213-233. 209 LiTERATuRE CiTED He, H.S.; Mladenoff, D.J.; Gustafson, E.J. 2002. Study of landscape change under forest harvesting and climate warming-induced fire disturbance. Forest Ecology and Management. 155(1): 257-270. Headwaters Economics. 2011. Economic profile system-human dimensions toolkit. Available at http://headwaterseconomics.org/tools/eps-hdt. (Accessed July 11, 2011). Hellmann, J.J.; Byers, J.E.; Bierwagen, B.G.; Dukes, J.S. 2008. Five potential consequences of climate change for invasive species. Conservation Biology. 22(3): 534-543. Hessl, A.; Pederson, N. 2013. Hemlock legacy project (HeLP): a paleoecological requiem for eastern hemlock. Progress in Physical Geography. 37(1): 114-129. Hessl, A.; Spackman, S. 1995. Effects of fire on threatened and endangered plants: an annotated bibliography. Information and Technology Report. Boulder, CO: U.S. Department of the Interior, National Biological Service. 55 p. Available at http://www.fws. gov/fire/ifcc/esr/library/teplants.pdf. (Accessed September 24, 2014). Hessl, A.E.; Saladyga, T.; Schuler, T.; Clark, P.; Wixom, J. 2011. Fire history from three species on a central Appalachian ridgetop. Canadian Journal of Forest Research. 41(10): 2031-2039. Hewes, J.H.; Butler, B.J.; Liknes, G.C.; Nelson, M.D.; Snyder, S.A. 2014. Public and private forest ownership in the conterminous United States: distribution of six ownership types - geospatial dataset. Fort Collins, CO: Forest Service Research Data Archive. Available at http://dx.doi.org/10.2737/RDS-2014-0002. (Accessed December 1, 2014). 210 Hicke, J.A.; Allen, C.D.; Desai, A.R.; Dietze, M.C.; Hall, R.J.; Kashian, D.M.; Moore, D.; Raffa, K.F.; Sturrock, R.N.; Vogelmann, J. 2012. Effects of biotic disturbances on forest carbon cycling in the United States and Canada. Global Change Biology. 18(1): 7-34. Hix, D.M.; Pearcy, J.N. 1997. Forest ecosystems of the Marietta Unit, Wayne National Forest, southeastern Ohio: multifactor classification and analysis. Canadian Journal of Forest Research. 27(7): 1117-1131. Hodgkins, G.A.; Dudley, R.W. 200. Changes in the timing of winter-spring streamflows in eastern North America, 1913-2002. Geophysical Research Letters. 33(): L0402. Hoekstra, J.M.; Molnar, J.L.; Jennings, M.; Revenga, C.; Spalding, M.D. 2010. The atlas of global conservation. Berkeley, CA: University of California Press: 272 p. Holl, K.D. 2002. Long-term vegetation recovery on reclaimed coal surface mines in the eastern USA. Journal of Applied Ecology. 39(): 90-970. Holland, G.J.; Webster, P.J. 2007. Heightened tropical cyclone activity in the North Atlantic: Natural variability or climate trend? Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 35(180): 295-271. Holzmueller, E.J.; Gibson, D.J.; Suchecki, P.F. 2012. Accelerated succession following an intense wind storm in an oak-dominated forest. Forest Ecology and Management. 279: 141-14. Horsley, S.B.; Stout, S.L.; deCalesta, D.S. 2003. White-tailed deer impact on the vegetation dynamics of a northern hardwood forest. Ecological Applications. 13(1): 98-118. LiTERATuRE CiTED Horton, R.; Yohe, G.; Easterling, W.; Kates, D.R.W.; Ruth, M.; Sussman, E.; Whelchel, A.; Wolfe, D. 2013. National Climate Assessment, chapter 16: northeast. Draft for Public Comment (v.11 Jan 2013). Available at http://ncadac. globalchange.gov/download/NCAJan112013-publicreviewdraft-chap1-northeast.pdf. (Accessed August 22, 2014). Hosaka, M.; Nohara, D.; Kitoh, A. 2005. Changes in snow cover and snow water equivalent due to global warming simulated by a 20km-mesh global atmospheric model. Sola. 1(0): 93-9. Hurlbert, A.H.; Liang, Z. 2012. Spatiotemporal variation in avian migration phenology: citizen science reveals effects of climate change. PLoS ONE. 7(2): e312. Hutchinson, T.F.; Vankat, J.L. 1998. Landscape structure and spread of the exotic shrub Lonicera maackii (Amur honeysuckle) in Southwestern Ohio forests. The American Midland Naturalist. 139(2): 383-390. Ibáñez, I.; Clark, J.S.; Dietz, M.C. 2008. Evaluating the sources of potential migrant species: implications under climate change. Ecological Applications. 18(7): 14-178. Ibáñez, I.; Clark, J.S.; Dietze, M.C.; Feeley, K.; Hersh, M.; LaDeau, S.; McBride, A.; Welch, N.E.; Wolosin, M.S. 200. Predicting biodiversity change: outside the climate envelope, beyond the species-area curve. Ecology. 87(8): 189-190. Intergovermental Panel on Climate Change [IPCC]. 2000. Special report on emission scenarios. In: Nakicenovic, N.; Swart, R., eds. Cambridge, UK: Cambridge University Press. 570 p. Available at http://www.grida.no/publications/ other/ipcc_sr/?src=/climate/ipcc/emission/. (Accessed August 21, 2014). Intergovermental Panel on Climate Change. [IPCC]. 2007. Climate change 2007: synthesis report. Contribution of Working Groups I, II, and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Core Writing Team, Pachauri, R.K. and Reisinger, A. (eds.)]. Geneva, Switzerland. 104 p. Available at http://www.ipcc.ch/publications_ and_data/publications_ipcc_fourth_assessment_ report_synthesis_report.htm. (Accessed August 21, 2014). Intergovermental Panel on Climate Change. [IPCC]. 2012. Managing the risks of extreme events and disasters to advance climate change adaptation. In: Field, C.B.; Barros, V.; Stocker, T.F.; Qin, D.; Dokken, D.J.; Ebi, K.L.; Mastrandrea, M.D.; Mach, K.J.; Plattner, G.-K.; Allen, S.K.; Tignor, M.; and Midgley, P. M., eds. A special report of Working Groups I and II of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York, NY: Cambridge University Press. 582 p. Irwin, E.G.; Bockstael, N.E. 2007. The evolution of urban sprawl: evidence of spatial heterogeneity and increasing land fragmentation. Proceedings of the National Academy of Sciences. 104(52): 2072-2077. Iverson, L.R.; Dale, M.E.; Scott, C.T.; Prasad, A. 1997. A GIS-derived integrated moisture index to predict forest composition and productivity of Ohio forests (USA). Landscape Ecology. 12(5): 331-348. Iverson, L.R.; Prasad, A.M. 1998. Predicting abundance of 80 tree species following climate change in the eastern United States. Ecological Monographs. 8(4): 45-485. Iverson, L.R.; Prasad, A.M.; Matthews, S. 2008a. Modeling potential climate change impacts on the trees of the northeastern United States. Mitigation and Adaptation Strategies for Global Change. 13(5): 487-51. 211 LiTERATuRE CiTED Iverson, L.R.; Prasad, A.M.; Matthews, S.N.; Peters, M. 2008b. Estimating potential habitat for 134 eastern US tree species under six climate scenarios. Forest Ecology and Management. 254(3): 390-40. Jeong, S.-J.; Ho, C.-H.; Gim, H.-J.; Brown, M.E. 2011. Phenology shifts at start vs. end of growing season in temperate vegetation over the Northern Hemisphere for the period 19822008. Global Change Biology. 17(7): 2385-2399. Iverson, L.R.; Schwartz, M.W.; Prasad, A.M. 2004a. How fast and far might tree species migrate in the eastern United States due to climate change? Global Ecology and Biogeography. 13(3): 209-219. Jonas, S.Z.; Xi, W.; Waldron, J.D.; Coulson, R.N. 2012. The impacts of hemlock decline and the ecological considerations for forest restoration following Hemlock Woolly Adelgid outbreaks. Tree and Forestry Science and Biotechnology. (1): 27-33. Iverson, L.R.; Schwartz, M.W.; Prasad, A.M. 2004b. Potential colonization of new available tree species habitat under climate change: an analysis for five eastern US species. Landscape Ecology. 19: 787-799. Jacob, D.J.; Winner, D.A. 2009. Effect of climate change on air quality. Atmospheric Environment. 43(1): 51-3. Jacobs, D.F.; Dalgleish, H.J.; Nelson, C.D. 2013. A conceptual framework for restoration of threatened plants: the effective model of American chestnut (Castanea dentata) reintroduction. New Phytologist. 197(2): 378-393. Jacobson, G.L., Jr.; Webb, T., III; Grimm, E.C. 1987. Patterns and rates of vegetation change during the deglaciation of eastern North America. In: The geology of North America. Vol. K-3: North America and adjacent oceans during the last deglaciation. Boulder, CO: The Geological Society of America: 277-288. Janetos, A.; Hansen, L.; Inouye, D.W.; Kelly, B.P.; Meyerson, L.; Peterson, B.; Shaw, R. 2008. Biodiversity. In: The effects of climate change on agriculture, land resources, water resources, and biodiversity in the United States. Washington, DC: U.S. Climate Change Science Program and the Subcommittee on Global Change Research. 151-181. 212 Jones, J.A.; Creed, I.F.; Hatcher, K.L.; Warren, R.J.; Adams, M.B.; Benson, M.H.; Boose, E.; Brown, W.A.; Campbell, J.L.; Covich, A. 2012. Ecosystem processes and human influences regulate streamflow response to climate change at long-term ecological research sites. BioScience. 2(4): 390-404. Joyce, J.M.; Scott, M.S. 2005. An assessment of Maryland’s vulnerability to flood damage. Baltimore, MD: Maryland Department of the Environment. 87 p. Available at http://www.esrgc. org/pdf/hazus/An%20Assessment%20of%20Mar yland’s%20Vulnerability%20to%20Flooding.pdf. (Accessed September 11, 2014). Julien, Y.; Sobrino, J.A. 2009. Global land surface phenology trends from GIMMS database. International Journal of Remote Sensing. 30(13): 3495-3513. Jump, A.S.; Peñuelas, J. 2005. Running to stand still: adaptation and the response of plants to rapid climate change. Ecology Letters. 8(9): 1010-1020. Kapnick, S.B.; Delworth, T.L. 2013. Controls of global snow under a changed climate. Journal of Climate. 2: 5537-552. Karl, T.R.; Knight, R.W.; Easterling, D.R.; Quayle, R.G. 199. Indices of climate change for the United States. Bulletin of the American Meteorological Society. 77(2): 279-292. LiTERATuRE CiTED Karl, T.R.; Melillo, J.M.; Peterson, T.C., eds. 2009. Global climate change impacts in the United States: a state of knowledge report from the U.S. Global Change Research Program. New York, NY: Cambridge University Press. 189 p. Kasey, P. 2012. Which county leads WV’s Marcellus production? The State Journal. Charleston, WV: WorldNow and WVSTATE. Available at http://www.statejournal.com/ story/19853093/which-county-leads-wvsmarcellus-production. (Accessed September 11, 2014). Kauserud, H.; Stige, L.C.; Vik, J.O.; Økland, R.H.; Høiland, K.; Stenseth, N.C. 2008. Mushroom fruiting and climate change. Proceedings of the National Academy of Sciences. 105(10): 3811-3814. Keyser, T.L.; Brown, P.M. 2014. Long-term response of yellow-poplar to thinning in the southern Appalachian Mountains. Forest Ecology and Management. 312(0): 148-153. Kitajima, K.; Fenner, M. 2000. Ecology of seedling regeneration. In: Fenner, M., ed. Seeds: the ecology of regeneration in plant communities. Wallingford, UK: CABI: 331-359. Kliejunas, J.T. 2011. A risk assessment of climate change and the impact of forest diseases on forest ecosystems in the Western United States and Canada. Gen. Tech. Rep. PSW-GTR-23. Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Station. 70 p. Available at http://www.fs.fed. us/psw/publications/documents/psw_gtr23/psw_ gtr23.pdf. (Accessed August 21, 2014). Kling, G.W.; Hayhoe, K.; Johnson, L.B.; Magnuson, J.J.; Polasky, S.; Robinson, S.K.; Shuter, B.J.; Wander, M.M.; Wuebbles, D.J.; Zak, D.R.; Lindroth, R.L.; Moser, S.C.; M.L.Wilson. 2003. Confronting climate change in the Great Lakes region: impacts on our communities and ecosystems. Cambridge, MA: Union of Concerned Scientists; Ecological Society of America. Available at http://ucsusa.org/assets/ documents/global_warming/greatlakes_final.pdf. (Accessed November 23, 2013). Knapp, A.K.; Beier, C.; Briske, D.D.; Classen, A.T.; Luo, Y.; Reichstein, M.; Smith, M.D.; Smith, S.D.; Bell, J.E.; Fay, P.A. 2008. Consequences of more extreme precipitation regimes for terrestrial ecosystems. BioScience. 58(9): 811-821. Knicker, H. 2007. How does fire affect the nature and stability of soil organic nitrogen and carbon? A review. Biogeochemistry. 85(1): 91-118. Knight, M. 2012. Dominion power outages third-largest after hurricanes Isabel and Irene. Available at http://wtkr.com/2012/07/01/ dominion-power-outages-third-largest-afterhurricanes-isabel-and-irene/. (Accessed January 28, 2014). Knight, T.M.; Dunn, J.L.; Smith, L.A.; Davis, J.; Kalisz, S. 2009. Deer facilitate invasive plant success in a Pennsylvania forest understory. Natural Areas Journal. 29(2): 110-11. Kochenderfer, J.N.; Adams, M.B.; Miller, G.W.; Helvey, J.D. 2007. Factors affecting large peakflows on Appalachian watersheds: lessons from the Fernow Experimental Forest. Res. Pap. NRS-3. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 24 p. 213 LiTERATuRE CiTED Koss, W.J.; Owensby, J.R.; Steurer, P.M.; Ezell, D.S. 1988. Freeze/frost data. Climatography of the U.S. Asheville, NC: National Climatic Data Center. 20 p. Kovacs, K.F.; Haight, R.G.; McCullough, D.G.; Mercader, R.J.; Siegert, N.W.; Liebhold, A.M. 2010. Cost of potential emerald ash borer damage in U.S. communities, 2009-2019. Ecological Economics. 9(3): 59-578. Kovarik, J.L. 2013. Geologic management of cave and karst resources on national forest system lands. Proceedings from the National Cave and Karst Management Symposium; Carlsbad, NM; November 4-7, 2013. Available at http:// scholarcommons.usf.edu/cgi/viewcontent.cgi?a rticle=1013&context=nckms_2013. (Accessed January 30, 2015). Krist, F.J.; Sapio, F.J.; Tkacz, B.M. 2007. Mapping risk from forest insects and diseases. Available at http://www.fs.fed.us/foresthealth/technology/ pdfs/FHTET2007-0_RiskMap.pdf. (Accessed September 11, 2014). Kunkel, K.; Bromirski, P.; Brooks, H.; Cavazos, T.; Douglas, A.; Easterling, D.; Emanuel, K.; Groisman, P.Y.; Holland, G.; Knutson, T. 2008. Observed changes in weather and climate extremes. In: Weather and climate extremes in a changing climate. Regions of focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. Washington, DC: U.S. Climate Change Science Program and the Subcommittee on Global Change Research: 35-80. 214 Kunkel, K.E.; Karl, T.R.; Brooks, H.; Kossin, J.; Lawrimore, J.H.; Arndt, D.; Bosart, L.; Changnon, D.; Cutter, S.L.; Doesken, N.; Emanuel, K.; Groisman, P.Y.; Katz, R.W.; Knutson, T.; O’Brien, J.; Paciorek, C.J.; Peterson, T.C.; Redmond, K.; Robinson, D.; Trapp, J.; Vose, R.; Weaver, S.; Wehner, M.; Wolter, K.; Wuebbles, D. 2013a. Monitoring and understanding trends in extreme storms: state of knowledge. Bulletin of the American Meteorological Society. 94(4): 499-514. Kunkel, K.E.; Liang, X.-Z.; Zhu, J.; Lin, Y. 200. Can CGCMs simulate the twentieth-century “warming hole” in the central United States? Journal of Climate. 19(17): 4137-4153. Kunkel, K.E.; Palecki, M.A.; Ensor, L.; Easterling, D.; Hubbard, K.G.; Robinson, D.; Redmond, K. 2009a. Trends in twentieth-century U.S. extreme snowfall seasons. Journal of Climate. 22(23): 204-21. Kunkel, K.E.; Palecki, M.; Ensor, L.; Hubbard, K.G.; Robinson, D.; Redmond, K.; Easterling, D. 2009b. Trends in twentieth-century U.S. snowfall using a quality-controlled dataset. Journal of Atmospheric and Oceanic Technology. 2(1): 33-44. Kunkel, K.E.; Stevens, L.E.; Stevens, S.E.; Sun, L.; Janssen, E.; Wuebbles, D.; Hilberg, S.D.; Timlin, M.S.; Stoecker, L.; Westcott, N.E.; Dobson, J.G. 2013b. Regional climate trends and scenarios for the U.S. National Climate Assessment. Part 3 - climate of the Midwest U.S. Washington, DC: U.S. Department of Commerce, National Oceanic and Atmospheric Administration. 102 p. Available at http://scenarios.globalchange. gov/sites/default/files/NOAA_NESDIS_Tech_ Report_142-3-Climate_of_the_Midwest_U.S_ 0.pdf. (Accessed August 21, 2014). LiTERATuRE CiTED Kunkel, K.E.; Stevens, L.E.; Stevens, S.E.; Sun, L.; Janssen, E.; Wuebbles, D.; Rennells, J.; DeGaetano, A.; Dobson, J.G. 2013c. Regional climate trends and scenarios for the U.S. National Climate Assessment. Part 1 - climate of the Northeast U.S. 79 p. Available at http:// scenarios.globalchange.gov/sites/default/files/ NOAA_NESDIS_Tech_Report_142-1-Climate_ of_the_Northeast_U.S_1.pdf. (Accessed August 21, 2014). Kunkel, K.E.; Westcott, N.E.; Kristovich, D.A.R. 2002. Assessment of potential effects of climate change on heavy lake-effect snowstorms near Lake Erie. Journal of Great Lakes Research. 28(4): 521-53. Kurtz, C.M. 2013. An assessment of invasive plant species monitored by the Northern Research Station Forest Inventory and Analysis program, 2005 through 2010. Gen. Tech. Rep. NRS-109. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 70 p. Landscape Change Research Group. 2014. Climate change atlas. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. Available at www.nrs. fs.fed.us/atlas. (Accessed September 8, 2014). Le Quéré, C.; Raupach, M.R.; Canadell, J.G.; Marland, G. 2009. Trends in the sources and sinks of carbon dioxide. Nature Geoscience. 2: 831-83. Lee, C.C. 2012. Utilizing synoptic climatological methods to assess the impacts of climate change on future tornado-favorable environments. Natural Hazards. 2(2): 325-343. Leibensperger, E.; Mickley, L.; Jacob, D.; Chen, W.; Seinfeld, J.; Nenes, A.; Adams, P.; Streets, D.; Kumar, N.; Rind, D. 2012. Climatic effects of 1950-2050 changes in US anthropogenic aerosols - Part 2: climate response. Atmospheric Chemistry and Physics. 12(7): 3349-332. Lemieux, C.J.; Scott, D.J. 2005. Climate change, biodiversity conservation and protected area planning in Canada. The Canadian Geographer/ Le Géographe canadien. 49(4): 384-397. Lemke, P.; Ren, J.; Alley, R.B.; Allison, I.; Carrasco, J.; Flato, G.; Fujii, Y.; Kaser, G.; Mote, P.; Thomas, R.H.; Zhang, T. 2007. Observations: changes in snow, ice and frozen ground. In: Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K.B.; Tignor, M.; Miller, H.L., eds. Climate change 2007: the physical science basis; summary for policymakers, technical summary and frequently asked questions. Part of the Working Group I contribution to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York, NY: Cambridge University Press: 337-383. Available at http:// www.ipcc.ch/pdf/assessment-report/ar4/wg1/ar4wg1-chapter4.pdf. (Accessed September 24, 2014). Lenihan, J.M.; Bachelet, D.; Neilson, R.P.; Drapek, R. 2008. Simulated response of conterminous United States ecosystems to climate change at different levels of fire suppression, CO2 emission rate, and growth response to CO2 . Global and Planetary Change. 4(1): 1-25. Likens, G.E.; Bormann, F.H. 1995. Biogeochemistry of a forested ecosystem. New York, NY: Springer-Verlag New York Inc. Linderholm, H.W. 200. Growing season changes in the last century. Agricultural and Forest Meteorology. 137(1-2): 1-14. Lins, H.F.; Slack, J.R. 1999. Streamflow trends in the United States. Geophysical Research Letters. 2(2): 227-230. Lippincott, C.L. 2000. Effects of Imperata cylindrica (L.) Beauv. (Cogongrass) invasion on fire regime in Florida sandhill. Natural Areas Journal. 20(2): 140-149. 215 LiTERATuRE CiTED Lister, T.W.; Perdue, J. 2013. Maryland’s forest resources, 2012. Res. Note NRS-191. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 4 p. Available at http://www.fs.fed.us/nrs/pubs/rn/ rn_nrs191.pdf. Liu, J.; Liu, S.; Loveland, T.R. 200. Temporal evolution of carbon budgets of the Appalachian forests in the U.S. from 1972 to 2000. Forest Ecology and Management. 222: 191-201. Liu, X.; Zhang, X.; Zhang, M. 2008. Major factors influencing the efficacy of vegetated buffers on sediment trapping: a review and analysis. Journal of Environmental Quality. 37(5): 17-174. Liu, Y.; Stanturf, J.; Goodrick, S. 2010. Trends in global wildfire potential in a changing climate. Forest Ecology and Management. 259(4): 85-97. Loeb, S.C.; Winters, E.A. 2013. Indiana bat summer maternity distribution: effects of current and future climates. Ecology and Evolution. 3(1): 103-114. Loehle, C. 1988. Tree life history strategies: the role of defenses. Canadian Journal of Forest Research. 18(2): 209-222. Logan, J.A.; Regniere, J.; Powell, J.A. 2003. Assessing the impacts of global warming on forest pest dynamics. Frontiers in Ecology and the Environment. 1(3): 130-137. Long, R.P.; Horsley, S.B.; Hall, T.J. 2013. Longterm impact of liming on growth and vigor of northern hardwoods. Canadian Journal of Forest Research. 41(): 1295-1307. Loomis, J.B.; Crespi, J. 1999. Estimated effects of climate change on selected outdoor recreation activities in the United States. In: Mendelsohn, R.; Neumann, J.E. The impact of climate change on the United States economy. Cambridge, UK: Cambridge University Press: 289-314. 21 Lovett, G.M.; Canham, C.D.; Arthur, M.A.; Weathers, K.C.; Fitzhugh, R.D. 200. Forest ecosystem responses to exotic pests and pathogens in eastern North America. BioScience. 5(5): 395-405. Madsen, T.; Willcox, N. 2012. When it rains, it pours: global warming and the increase in extreme precipitation from 1948 to 2011. Trenton, NJ: Environment America Research & Policy Center. 47 p. Maherali, H.; Moura, C.F.; Caldeira, M.C.; Willson, C.J.; Jackson, R.B. 200. Functional coordination between leaf gas exchange and vulnerability to xylem cavitation in temperate forest trees. Plant, Cell & Environment. 29(4): 571-583. Malmsheimer, R.W.; Bowyer, J.L.; Fried, J.S.; Gee, E.; Izlar, R.L.; Miner, R.A.; Munn, I.A.; Oneil, E.; Stewart, W.C. 2011. Managing forests because carbon matters: integrating energy, products, and land management policy. Journal of Forestry. 109(Supplement 1): S7-S51. Mastrandrea, M.D.; Field, C.B.; Stocker, T.F.; Edenhofer, O.; Ebi, K.L.; Frame, D.J.; Held, H.; Kriegler, E.; Mach, K.J.; Matschoss, P.R.; Plattner, G.-K.; Yohe, G.W.; Zwiers, F.W. 2010. Guidance note for lead authors of the IPCC Fifth Assessment Report on consistent treatment of uncertainties. Intergovernmental Panel on Climate Change (IPCC). Available at http://www.ipcc.ch/activities/activities.shtml. (Accessed Feb. 28, 2011). Matthews, S.N.; Iverson, L.R.; Prasad, A.M.; Peters, M.P. 2011a. Changes in potential habitat of 147 North American breeding bird species in response to redistribution of trees and climate following predicted climate change. Ecography. 34(): 933-945. LiTERATuRE CiTED Matthews, S.N.; Iverson, L.R.; Prasad, A.M.; Peters, M.P.; Rodewald, P.G. 2011b. Modifying climate change habitat models using tree speciesspecific assessments of model uncertainty and life history-factors. Forest Ecology and Management. 22(8): 140-1472. Maurer, E.; Wood, A.; Adam, J.; Lettenmaier, D.; Nijssen, B. 2002. A long-term hydrologically based dataset of land surface fluxes and states for the conterminous United States. Journal of Climate. 15(22): 3237-3251. McConnell, M.E.; Balci, Y. 2014. Phytophthora cinnamomi as a contributor to white oak decline in Mid-Atlantic United States forests. Plant Disease 98(3): 319-327. McCullough, D.G.; Werner, R.A.; Neumann, D. 1998. Fire and insects in northern and boreal forest ecosystems of North America. Annual Review of Entomology. 43(1): 107-127. McEwan, R.W.; Brecha, R.J.; Geiger, D.R.; John, G.P. 2011. Flowering phenology change and climate warming in southwestern Ohio. Plant Ecology. 212(1): 55-1. McLachlan, J.S.; Clark, J.S.; Manos, P.S. 2005. Molecular indicators of tree migration capacity under rapid climate change. Ecology. 8(8): 2088-2098. McLaughlin, S.; Andersen, C.; Edwards, N.; Roy, W.; Layton, P. 1990. Seasonal patterns of photosynthesis and respiration of red spruce saplings from two elevations in declining southern Appalachian stands. Canadian Journal of Forest Research. 20(5): 485-495. McLaughlin, S.B.; Kohut, R.J. 1992. The effects of atmospheric deposition and ozone on carbon allocation and associated physiological processes in red spruce. In: Eagar, C.; Adams, M.B., eds. Ecology and decline of red spruce in the eastern United States. New York, NY: Springer-Verlag: 338-382. McMahon, S.M.; Parker, G.G.; Miller, D.R. 2010. Evidence for a recent increase in forest growth. Proceedings of the National Academy of Sciences. 107(8): 311-315. McNab, W.; Cleland, D.; Freeouf, J.; Keys, J., Jr.; Nowacki, G.; Carpenter, C. 2007. Description of ecological subregions: sections of the conterminous United States. Gen. Tech. Rep. WO-7B. Washington, DC: U.S. Department of Agriculture, Forest Service. Available at http:// www.edc.uri.edu/aTmT-dSS/ReportForecast/ SectionDescriptions.pdf. (Accessed December , 2011). McNab, W.H.; Avers, P.E. 1994. Ecological subregions of the United States: section descriptions. Administrative Publication WO-WSA-5. Washington, DC: U.S. Department of Agriculture, Forest Service, Ecosystem Management. 27 p. Available at http://www.fs.fed.us/land/pubs/ecoregions/. (Accessed September 22, 2014). McNulty, S.G.; Boggs, J.L. 2010. A conceptual framework: redefining forest soil’s critical acid loads under a changing climate. Environmental Pollution. 158(): 2053-2058. Maryland Department of Natural Resources [MDNR]. 2010. Maryland forest resource assessment. Annapolis, MD: Maryland Department of Natural Resources, Forest Service. 90 p. Maryland Department of Natural Resources [MDNR]. 2014. Forest conservation and management program. Annapolis, MD: Maryland Department of Natural Resources, Forest Service. Available at http://www.dnr.state. md.us/forests/programapps/fcmp.html. (Accessed January 14, 2014). Meehl, G.A.; Arblaster, J.M.; Branstator, G. 2012. Mechanisms contributing to the warming hole and the consequent US east-west differential of heat extremes. Journal of Climate. 25(18): 394-408. 217 LiTERATuRE CiTED Meehl, G.A.; Tebaldi, C. 2004. More intense, more frequent, and longer lasting heat waves in the 21st century. Science. 305(58): 994-997. Melillo, J.M.; Richmond, T.C.; Yohe, G.W., eds. 2014. Climate change impacts in the United States: the third National Climate Assessment. Washington, DC: U.S. Global Change Research Program. 841 p. Available at http://nca2014.globalchange.gov/downloads. (Accessed August 22, 2014). Mendelsohn, R.; Neumann, J.E. 2004. The impact of climate change on the United States economy. Cambridge, UK: Cambridge University Press. 344 p. Merkle, S.; Andrade, G.; Nairn, C.; Powell, W.; Maynard, C. 2007. Restoration of threatened species: a noble cause for transgenic trees. Tree Genetics & Genomes. 3(2): 111-118. Mickley, L.; Jacob, D.; Field, B.; Rind, D. 2004. Effects of future climate change on regional air pollution episodes in the United States. Geophysical Research Letters. 31: L24103. Miles, P.D.; Brand, G.J.; Alerich, C.L.; Bednar, L.F.; Woudenberg, S.W.; Glover, J.F.; Ezell, E.N. 200. The Forest Inventory and Analysis database description and users manual version 1.0. Gen. Tech. Rep. NC-218. St. Paul, MN: U.S. Department of Agriculture, Forest Service, North Central Research Station. 130 p. Mladenoff, D.J. 2004. LANDIS and forest landscape models. Ecological Modelling. 180(1): 7-19. Mogil, H.M.; Seaman, K.L. 2009. The climate and weather of Delaware, Maryland, and Washington, D.C. Philadelphia, PA: Taylor & Francis Group. Available at http://www. weatherwise.org/Archives/Back%20Issues/2009/ July-August%202009/full-mogil.html. (Accessed September 9, 2014). 218 Moore, M.V.; Pace, M.L.; Mather, J.R.; Murdoch, P.S.; Howarth, R.W.; Folt, C.L.; Chen, C.Y.; Hemond, H.F.; Flebbe, P.A.; Driscoll, C.T. 1997. Potential effects of climate change on freshwater ecosystems of the New England/ Mid-Atlantic region. Hydrological Processes. 11(8): 925-947. Morin, R.S.; Liebhold, A.M.; Tobin, P.C.; Gottschalk, K.W.; Luzader, E. 2007. Spread of beech bark disease in the eastern United States and its relationship to regional forest composition. Canadian Journal of Forest Research. 37(4): 72-73. Moritz, M.A.; Parisien, M.A.; Batllori, E.; Krawchuk, M.A.; Van Dorn, J.; Ganz, D.J.; Hayhoe, K. 2012. Climate change and disruptions to global fire activity. Ecosphere. 3: 49. Morris, D.; Walls, M.A. 2009. Climate change and outdoor recreation resources. Washington, DC: Resources for the Future. Moscow Forest Sciences Laboratory. 2014. Climate estimates and plant-climate relationships: current climate data. Moscow, ID: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. Available at http://forest.moscowfsl.wsu.edu/. (Accessed September 19, 2014). Munger, G.T. 2001. Alliaria petiolata. Fire effects information system. [Online]. Missoula, MT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory. Available at http://www.fs.fed.us/ database/feis/plants/forb/allpet/introductory.html. (Accessed September 22, 2014). Munger, G.T. 2002. Pueraria montana var. lobata. Fire effects information system. [Online]. Missoula, MT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory. Available at http://www.fs.fed.us/database/feis/plants/vine/ puemonl/introductory.html. (Accessed September 22, 2014). LiTERATuRE CiTED Murphy, H.T.; VanDerWal, J.; Lovett-Doust, J. 2010. Signatures of range expansion and erosion in eastern North American trees. Ecology Letters. 13(10): 1233-1244. National Interagency Fire Center Statistics. 2013. Lightning vs. human caused fires and acres (stats reported from 2001). [Online]. Boise, ID. Available at http://www.nifc.gov/fireInfo/ fireInfo_stats_lightng.html. (Accessed September 8, 2014). National Oceanic and Atmospheric Administration [NOAA]. 2008. Top ten breakthroughs: the first climate model. Washington, DC. Available at http://celebrating200years.noaa.gov/ breakthroughs/climate_model/welcome.html. National Oceanic and Atmospheric Administration [NOAA]. 2012. State of the climate: national overview for annual 2012. National Climatic Data Center. Available at http://www.ncdc.noaa. gov/sotc/national/2012/13. (Accessed May 8, 2013). National Oceanic and Atmospheric Administration [NOAA]. 2013a. Enhanced Fujita scale (EF scale). Norman, OK: National Weather Service. Available at http://www.spc.noaa.gov/efscale/. (Accessed June 19, 2013). National Oceanic and Atmospheric Administration [NOAA]. 2013b. Historic “derecho” severe wind event storms across the Midwest and Mid-Atlantic on June 29, 2012. Blacksburg, VA: National Weather Service, Weather Forecast Office. Available at http://www.erh.noaa.gov/ rnk/events/2012/Jun29_derecho/summary.php. (Accessed October 3, 2013). National Oceanic and Atmospheric Administration [NOAA]. 2013c. The Ozarks / Ohio Valley derecho of April 2011. Storm Prediction Center. Available at http://www.spc.noaa.gov/misc/ AbtDerechos/casepages/apr202011page.htm. (Accessed August 21, 2014). National Oceanic and Atmospheric Administration [NOAA]. 2014a. Billion-dollar weather/climate disasters. National Climatic Data Center. Available at http://www.ncdc.noaa.gov/billions/ events. (Accessed September 10, 2014). National Oceanic and Atmospheric Administration [NOAA]. 2014b. Climate at a glance. National Climatic Data Center. Available at http://www.ncdc.noaa.gov/cag/time-series/global. (Accessed January 27, 2014). National Oceanic and Atmospheric Administration [NOAA]. 2014c. Climate maps of the United States (1961-1990). National Climatic Data Center. Available at http://cdo.ncdc.noaa.gov/cgibin/climaps/climaps.pl?directive=quick_results& subrnum=&pop=YES. (Accessed September 8, 2014). National Oceanic and Atmospheric Administration [NOAA]. 2014d. National temperature and precipitation maps. National Climatic Data Center. Available at http://www.ncdc.noaa.gov/ temp-and-precip/maps.php. (Accessed August 21, 2014). National Research Council [NRC]. 2007. Environmental impacts of wind-energy projects. National Research Council, Committee on Environmental Impacts of Wind-Energy Projects. 394 p. Available at http://www.nap.edu/catalog/11935.html. (Accessed September 22, 2014). National Weather Service. 2012. Storm Prediction Center. Available at http://www.spc.noaa.gov/ wcm/. (Accessed April 1, 2012). Natural Resources Conservation Service [NRCS]. 200. Land resource regions and major land resource areas of the United States, the Carribbean, and the Pacific Basin. Agric. Handb. 29. 71 p. Available at http://www. nrcs.usda.gov/Internet/FSE_DOCUMENTS/ nrcs142p2_051845.pdf. (Accessed September 9, 2014). 219 LiTERATuRE CiTED Natural Resources Conservation Service [NRCS]. 2014a. NJ biology technical note: southern pine beetle and forest management. Available at http://www.nrcs.usda.gov/Internet/ FSE_DOCUMENTS/nrcs141p2_018399.pdf. (Accessed January 14, 2014). Natural Resources Conservation Service [NRCS]. 2014b. PLANTS database. [Online]. National Plant Data Team. Available at http://plants.usda. gov/java/. (Accessed September 9, 2014). The Nature Conservancy. 2003. The Cumberlands and Southern Ridge & Valley Ecoregion: a plan for biodiversity conservation. Arlington, VA. 232 p. Available at http://www. conservationgateway.org/ConservationPlanning/ SettingPriorities/EcoregionalReports/Documents/ CSRVPlan.pdf. (Accessed August 21, 2014). NatureServe. 2011. NatureServe Explorer. Arlington, VA. Available at http://www. natureserve.org/explorer/classeco.htm. (Accessed September 11, 2014). NatureServe. 2013. NatureServe Explorer: an online encyclopedia of life. Arlington, VA. Available at http://www.natureserve.org/explorer. (Accessed September 1, 2013). Nearing, M. 2001. Potential changes in rainfall erosivity in the US with climate change during the 21st century. Journal of Soil and Water Conservation. 5(3): 229-232. Nearing, M.; Pruski, F.; O’Neal, M. 2004. Expected climate change impacts on soil erosion rates: a review. Journal of Soil and Water Conservation. 59(1): 43-50. Nebel, S.; Mills, A.; McCracken, J.D.; Taylor, P.D. 2010. Declines of aerial insectivores in North America follow a geographic gradient. Avian Conservation and Ecology. 5(2): 1. 220 Neff, R.; Chang, H.; Knight, C.G.; Najjar, R.G.; Yarnal, B.; Walker, H.A. 2000. Impact of climate variation and change on Mid-Atlantic region hydrology and water resources. Climate Research. 14(3): 207-218. Negley, T.L.; Eshleman, K.N. 200. Comparison of stormflow responses of surface-mined and forested watersheds in the Appalachian mountains, USA. Hydrological Processes. 20(1): 347-3483. Nicholls, S. 2012. Outdoor recreation and tourism. In: Winkler, J.; Hatfield, J.; Bidwell, D.; Brown, D., eds. U.S. National Climate Assessment Midwest Technical Input Report Great Lakes Integrated Sciences and Assessments (GLISA) Center. Available at http://www.glisa. umich.edu/docs/NCA/MTIT_RecTourism.pdf. (Accessed September 10, 2014). Nilsson, C.; Brown, R.L.; Jansson, R.; Merritt, D.M. 2010. The role of hydrochory in structuring riparian and wetland vegetation. Biological Reviews. 85(4): 837-858. Norby, R.; DeLucia, E.; Gielen, B.; Calfapietra, C.; Giardina, C.; King, J.; Ledford, J.; McCarthy, H.; Moore, D.; Ceulemans, R. 2005. Forest response to elevated CO2 is conserved across a broad range of productivity. Proceedings of the National Academy of Sciences. 102(50): 18052. Nowacki, G.; Carr, R.; Van Dyck, M. 2009. The current status of red spruce in the eastern United States: distribution, population trends, and environmental drivers. In: Rentch, J.S.; Schuler, T.M., eds. Proceedings from the conference on ecology and management of high-elevation forests in the central and southern Appalachian Mountains. Gen. Tech. Rep. NRS-P-4. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station: 140-12. LiTERATuRE CiTED Nowacki, G.J.; Abrams, M.D. 2008. The demise of fire and “mesophication” of forests in the eastern United States. Bioscience. 58(2): 123-138. O’Malley, M. 2011. Executive order 01.01.2011.11: the Marcellus shale safe drilling initiative. Annapolis, MD: State of Maryland, Executive Department. Available at http://www.governor. maryland.gov/executiveorders/01.01.2011.11.pdf. (Accessed September 8, 2014). Ockels, F.S.; Mielke, M.; Bonello, P. 2014. Sudden oak death: monitoring Phytophthora ramorum in the North-central United States. Columbus, OH: Ohio State University. Available at http://ohioline.osu.edu/sc195/020.html. (Accessed May 14, 2014). Ohio Department of Natural Resources [ODNR]. 1998. Physiographic regions of Ohio. [Pagesize map with text, scale 1:2,100,00]. Division of Geological Survey. 2 p. Available at http://www. epa.state.oh.us/portals/27/SIP/Nonattain/F2physiographic_regions_of_Ohio.pdf. (Accessed September 9, 2014). Ohio Department of Natural Resources [ODNR]. 2010a. Invasive plants of Ohio: tree-of-heaven. Fact sheet 17. Columbus, OH. Available at http://ohiodnr.gov/portals/0/pdfs/invasives/ 17FactsheetTreeofHeaven.pdf. (Accessed September 18, 2014). Ohio Department of Natural Resources [ODNR]. 2005. Glacial map of Ohio. Division of Geological Survey. Available at http://geosurvey. ohiodnr.gov/portals/geosurvey/PDFs/Glacial/ glacial.pdf. (Accessed September 12, 2014). Ohio Department of Natural Resources [ODNR]. 2010b. Ohio’s statewide forest resource assessment. Columbus, OH: Division of Forestry. 188 p. Ohio Department of Natural Resources [ODNR]. 2013. Soil regions of Ohio. Division of Soil and Water Conservation. Available at http://www2. ohiodnr.com/soilwater/soil-conservation/ohiosoil-regions. (Accessed August 21, 2014). Ohio Department of Natural Resources [ODNR]. 2014. Ohio forest tax law (OFTL). Division of Forestry. Available at http://forestry.ohiodnr. gov/oftl. (Accessed August 21, 2014). Ohio Emergency Management Agency. 2011. State of Ohio hazard mitigation plan. Available at http://ema.ohio.gov/Documents/ OhioMitigationPlan/2011/Section%202_ HIRA%20Part%201.pdf. (Accessed August 21, 2014). Olano, J.M.; Palmer, M.W. 2003. Stand dynamics of an Appalachian old-growth forest during a severe drought episode. Forest Ecology and Management. 174(1-3): 139-148. Ollinger, S.; Goodale, C.; Hayhoe, K.; Jenkins, J. 2008. Potential effects of climate change and rising CO2 on ecosystem processes in northeastern U.S. forests. Mitigation and Adaptation Strategies for Global Change. 13(5): 47-485. Pan, Y.; Birdsey, R.; Hom, J.; McCullough, K. 2009. Separating effects of changes in atmospheric composition, climate and land-use on carbon sequestration of US Mid-Atlantic temperate forests. Forest Ecology and Management. 259(2): 151-14. Pan, Z.; Arritt, R.W.; Takle, E.S.; Gutowski, W.J.; Anderson, C.J.; Segal, M. 2004. Altered hydrologic feedback in a warming climate introduces a warming hole. Geophysical Research Letters. 31(17). 221 LiTERATuRE CiTED Pardo, L.H.; Robin-Abbott, M.J.; Driscoll, C.T. 2011. Assessment of nitrogen deposition effects and empirical critical loads of nitrogen for ecoregions of the United States. Gen.Tech. Rep. NRS-80. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 291 p. Parmesan, C. 2007. Influences of species, latitudes and methodologies on estimates of phenological response to global warming. Global Change Biology. 13(9): 180-1872. Parmesan, C.; Yohe, G. 2003. A globally coherent fingerprint of climate change impacts across natural systems. Nature. 421(918): 37-42. Pastor, J.; Post, W. 1985. Development of a linked forest productivity-soil process model. Oak Ridge National Laboratory, Environmental Sciences Division Publication No. 2455. Springfield, VA: U.S. Department of Commerce. 17 p. Patterson, W. 200. The paleoecology of fire and oaks in eastern forests. In: Dickinson, M.B., ed. Fire in eastern oak forests: delivering science to land managers, proceedings of a conference; 2005 November 15-17; Columbus, OH. Gen. Tech. Rep. NRS-P-1. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station: 2-19. Pennsylvania Department of Conservation and Natural Resources. 2013. Forest health fact sheet: hemlock woolly adelgid. Available at http://www.dcnr.state.pa.us/cs/groups/public/ documents/document/dcnr_007179.pdf. (Accessed August 21, 2014). Pennsylvania Department of Environmental Protection [PDEP]. 2004. Pennsylvania field guide: common invasive plants in riparian areas. 29 p. Available at http://www.dep.state. pa.us/dep/deputate/watermgt/wc/subjects/ streamreleaf/Docs/Invasive%20Plants.pdf. (Accessed August 21, 2014). 222 Pennsylvania Department of Environmental Protection [PDEP]. 2012. Erosion and sediment pollution control program manual. Technical Guidance No. 33-2124-008. 53 p. Available at http://www.elibrary.dep.state.pa.us/dsweb/Get/ Document-88925/33-2134-008.pdf. (Accessed August 21, 2014). Perera, E.M.; Sanford, T.; White-Newsome, J.L.; Kalkstein, L.S.; Vanos, J.K.; Weir, K. 2012. Heat in the Heartland: 60 years of warming in the Midwest. Union of Concerned Scientists. 40 p. Available at http://www.ucsusa.org/ assets/documents/global_warming/Heat-in-theHeartland-Full-Report.pdf. (Accessed August 21, 2014). Perry, L.B.; Konrad, C.E. 200. Relationships between NW flow snowfall and topography in the southern Appalachians, USA. Climate Research. 32(1): 35. Peterjohn, W.T.; Adams, M.B.; Gilliam, F.S. 199. Symptoms of nitrogen saturation in two central Appalachian hardwood forest ecosystems. Biogeochemistry. 35(3): 507-522. Peters, R.L.; Darling, J.D. 1985. The greenhouse effect and nature reserves. BioScience. 35(11): 707-717. Peters, M.P.; Iverson, L.R.; Matthews, S.N.; Prasad, A.M. 2013. Wildfire hazard mapping: exploring site conditions in eastern US wildland-urban interfaces. International Journal of Wildland Fire. 22(5): 57-578. Peterson, C.J. 2000. Catastrophic wind damage to North American forests and the potential impact of climate change. The Science of the Total Environment. 22: 287-311. Peterson, C.J.; Krueger, L.M.; Royo, A.A.; Stark, S.; Carson, W.P. 2013. Disturbance size and severity covary in small and mid-size wind disturbances in Pennsylvania northern hardwoods forests. Forest Ecology and Management. 302: 273-279. LiTERATuRE CiTED Peterson, G.C.; Allen, R.; C.S.Holling. 1998. Ecological resilience, biodiversity, and scale. Ecosystems. 1(1): -18. Peterson, T.C.; Zhang, X.; Brunet-India, M.; Vázquez-Aguirre, J.L. 2008. Changes in North American extremes derived from daily weather data. Journal of Geophysical Research. 113(D7): D07113. Pitchford, J.L.; Wu, C.; Lin, L.S.; Petty, J.T.; Thomas, R.; Veselka, W.E.; Welsch, D.; Zegre, N.; Anderson, J.T. 2012. Climate change effects on hydrology and ecology of wetlands in the Mid-Atlantic highlands. Wetlands. 32: 21-33. Piva, R.J.; Cook, G.W. 2011. West Virginia timber industry: an assessment of timber product output and use, 2007. Gen. Tech. Rep. NRS-4. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 4 p. Poff, N.L. 1992. Regional hydrologic response to climate change: an ecological perspective. In: Firth, P.; Fisher, S., eds. Global climate change and freshwater ecosystems. New York, NY: Springer: 88-115. Polsky, C.; Allard, J.; Currit, N.; Crane, R.; Yarnal, B. 2000. The Mid-Atlantic region and its climate: past, present, and future. Climate Research. 14(3): 11-173. Potter, K.M.; Hargrove, W.W.; Koch, F.H. 2010. Predicting climate change extirpation risk for central and southern Appalachian forest tree species. In: Rentch, J.S.; Schuler, T.M., eds. Proceedings from the conference on ecology and management of high-elevation forests of the central and southern Appalachian Mountains. Gen. Tech. Rep. NRS-P-4. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station: 179-189. Primack, R.B.; Miller-Rushing, A.J. 2012. Uncovering, collecting, and analyzing records to investigate the ecological impacts of climate change: a template from Thoreau’s Concord. BioScience. 2(2): 170-181. Raffa, K.F.; Aukema, B.H.; Bentz, B.J.; Carroll, A.L.; Hicke, J.A.; Turner, M.G.; Romme, W.H. 2008. Cross-scale drivers of natural disturbances prone to anthropogenic amplification: the dynamics of bark beetle eruptions. BioScience. 58(): 501-517. Raupach, M.R.; Marland, G.; Ciais, P.; Le Quere, C.; Canadell, J.G.; Klepper, G.; Field, C.B. 2007. Global and regional drivers of accelerating CO2 emissions. Proceedings of the National Academy of Sciences. 104(24): 10288-10293. Pope, V.; Gallani, M.L.; Rowntree, P.R.; Stratton, R.A. 2000. The impact of new physical parameterizations in the Hadley Centre climate model: HadAM3. Climate Dynamics. 1: 123-14. Rebbeck, J. 2012. Fire management and woody invasive plants in oak ecosystems. In: Dey, D.C.; Stambaugh, M.C.; Clark, S.L.; Schweitzer, C. J., eds. Proceedings of the 4th fire in eastern forests conference. Gen. Tech. Rep. NRS-P-102. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station: 142-155. Portmann, R.W.; Solomon, S.; Hegerl, G.C. 2009. Spatial and seasonal patterns in climate change, temperatures, and precipitation across the United States. Proceedings of the National Academy of Sciences. 10(18): 7324-7329. Rennenberg, H.; Dannenmann, M.; Gessler, A.; Kreuzwieser, J.; Simon, J.; Papen, H. 2009. Nitrogen balance in forest soils: nutritional limitation of plants under climate change stresses. Plant Biology. 11(s1): 4-23. Post, W.M.; Pastor, J. 199. Linkages—an individual-based forest ecosystem model. Climatic Change. 34(2): 253-21. 223 LiTERATuRE CiTED Resources for the Future. 2012. A review of shale gas regulations by state. Washington, DC: Resources For the Future (RFF), Center for Energy Economics and Policy. Available at http://www.rff.org/centers/energy_economics_ and_policy/Pages/Shale_Maps.aspx. (Accessed August 21, 2014). Reusch, T.B.H.; Ehlers, A.; Hammerli, A.; Worm, B. 2005. Ecosystem recovery after climatic extremes enhanced by genotypic diversity. Proceedings of the National Academy of Sciences. 102(8): 282-2831. Richardson, K.; Steffen, W.; Schellnhuber, H.J.; Alcamo, J.; Barker, T.; Kammen, D.; Leemans, R.; Liverman, D.; Monasinghe, M.; OsmanElasha, B.; Stern, N.; Waever, O. 2009. Synthesis report from Climate change: global risks challenges and decisions. Copenhagen, Denmark: University of Copenhagen. Available at http://edepot.wur.nl/10197. (Accessed August 21, 2014). Richardson, R.B.; Loomis, J.B. 2004. Adaptive recreation planning and climate change: a contingent visitation approach. Ecological Economics. 50(1): 83-99. Riitters, K.H. 2011. Spatial patterns of land cover in the United States. A technical document supporting the Forest Service 2010 RPA assessment. Gen. Tech. Rep. SRS-13. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 4 p. Riitters, K.H.; Wickham, J.D.; O’Neill, R.V.; Jones, K.B.; Smith, E.R.; Coulston, J.W.; Wade, T.G.; Smith, J.H. 2002. Fragmentation of continental United States forests. Ecosystems. 5(8): 0815-0822. 224 Robertson, A.W.; Kushnir, Y.; Lall, U.; Nakamura, J. 2011. On the connection between lowfrequency modulation of large-scale weather regimes and springtime extreme flooding over the Midwest of the United States. 3th Annual climate diagnostics and prediction workshop. National Oceanographic and Atmospheric Administration: Science and Technology Infusion Climate Bulletin: 150-152. Available at http:// www.nws.noaa.gov/ost/climate/STIP/3CDPW/ 3cdpw-arobertson.pdf. (Accessed August 21, 2014). Robinson, E.D.; Trapp, R.J.; Baldwin, M.E. 2013. The geospatial and temporal distributions of severe thunderstorms from high-resolution dynamical downscaling. Journal of Applied Meteorology and Climatology. 52: 2147-211. Root, T.L.; Price, J.T.; Hall, K.R.; Schneider, S.H.; Rosenzweig, C.; Pounds, J.A. 2003. Fingerprints of global warming on wild animals and plants. Nature. 421(918): 57-0. Runyan. 2011. Economic impact of travel on West Virginia 2000-2010: detailed state and county estimates. Portland, OR: Dean Runyan and Associates. Available at http://www.wvcommerce.org/App_Media/ assets/doc/travelandrec/industry/marketing/ 2010%20Economic%20Impact.pdf. (Accessed August 21, 2014). Rustad, L.; Campbell, J.; Marion, G.; Norby, R.; Mitchell, M.; Hartley, A.; Cornelissen, J.; Gurevitch, J. 2001. A meta-analysis of the response of soil respiration, net nitrogen mineralization, and aboveground plant growth to experimental ecosystem warming. Oecologia. 12(4): 543-52. LiTERATuRE CiTED Ryan, M.G.; Archer, S.R.; Birdsey, R.; Dahm, C.; Heath, L.; Hicke, J.; Hollinger, D.; Huxman, T.; Okin, G.; Oren, R.; Randerson, J.; Schlesinger, W. 2008. Land resources. In: The effects of climate change on agriculture, land resources, water resources, and biodiversity in the United States. Washington, DC: U.S. Climate Change Science Program and the Subcommittee on Global Change Research: 75-120. Ryan, M.G.; Harmon, M.E.; Birdsey, R.A.; Giardina, C.P.; Heath, L.S.; Houghton, R.A.; Jackson, R.B.; McKinley, D.C.; Morrison, J.F.; Murray, B.C. 2010. A synthesis of the science on forests and carbon for US forests. Issues In Ecology: Ecological Society of America. 13: 1-1. Sasek, T.W.; Strain, B.R. 1988. Effects of carbon dioxide enrichment on the growth and morphology of kudzu (Pueraria lobata). Weed Science. 3: 28-3. Sasek, T.W.; Strain, B.R. 1989. Effects of carbon dioxide enrichment on the expansion and size of kudzu (Pueraria lobata) leaves. Weed Science. 37: 23-28. Sasek, T.W.; Strain, B.R. 1990. Implications of atmospheric CO2 enrichment and climatic change for the geographical distribution of two introduced vines in the USA. Climatic Change. 1(1): 31-51. Saunders, S.; Findlay, D.; Easley, T.; Spencer, T. 2012. Doubled trouble: more Midwestern extreme storms. Louisville, CO: Rocky Mountain Climate Organization; New York, NY: Natural Resources Defense Council. 42 p. Available at http://www.rockymountainclimate. org/images/Doubled%20Trouble.pdf. (Accessed August 21, 2014). Schaberg, P.G.; Tilley, J.W.; Hawley, G.J.; DeHayes, D.H.; Bailey, S.W. 200. Associations of calcium and aluminum with the growth and health of sugar maple trees in Vermont. Forest Ecology and Management. 223(1): 159-19. Scheller, R.M.; Mladenoff, D.J. 2008. Simulated effects of climate change, fragmentation, and inter-specific competition on tree species migration in northern Wisconsin, USA. Climate Research. 3(3): 191-202. Schiffer, M.B. 199. Formation processes of the archaeological record. Salt Lake City, UT: University of Utah Press. 448 p. Schlesinger, M.D.; Corser, J.D.; Perkins, K.A.; White, E.L. 2011. Vulnerability of at risk species to climate change in New York. Albany, NY: New York Natural Heritage Program. 7 p. Available at https://connect.natureserve.org/ sites/default/files/documents/ccvi_report_ny.pdf. (Accessed August 21, 2014). Schneider, K.; Culver, D.C. 2004. Estimating subterranean species richness using intensive sampling and rarefaction curves in a high density cave region in West Virginia. Journal of Cave and Karst Studies. (2): 39-45. Schuler, T.M.; Collins, R.J. 2002. Successional dynamics and restoration implications of a montane coniferous forest in the central Appalachians, USA. Natural Areas Journal. 22(2): 88-98. Schuler, T.M.; Gillespie, A.R. 2000. Temporal patterns of woody species diversity in a central Appalachian forest from 1856 to 1997. Journal of the Torrey Botanical Society. 127: 149-11. Schwartz, M.D.; Ault, T.R.; Betancourt, J.L. 2013. Spring onset variations and trends in the continental United States: past and regional assessment using temperature-based indices. International Journal of Climatology. 33(13): 2917-2922. Schwartz, M.D.; R.Ahas; Aasa, A. 200a. Onset of spring starting earlier across the Northern Hemisphere. Global Change Biology. 12(2): 343-351. 225 LiTERATuRE CiTED Schwartz, M.W.; Iverson, L.R.; Prasad, A.M.; Matthews, S.N.; O’Connor, R.J. 200b. Predicting extinctions as a result of climate change. Ecology. 87(7): 111-115. Sisk, T.D.; Haddad, N.M.; Ehrlich, P.R. 1997. Bird assemblages in patchy woodlands: modeling the effects of edge and matrix habitats. Ecological Applications. 7(4):1170-1180. Seidel, T.M.; Weihrauch, D.M.; Kimball, K.D.; Pszenny, A.A.; Soboleski, R.; Crete, E.; Murray, G. 2009. Evidence of climate change declines with elevation based on temperature and snow records from 1930s to 2006 on Mount Washington, New Hampshire, USA. Arctic, Antarctic, and Alpine Research. 41(3): 32-372. Smith, D.L.; Johnson, L. 2004. Vegetationmediated changes in microclimate reduce soil respiration as woodlands expand into grasslands. Ecology. 85(12): 3348-331. Shang, B.Z.; He, H.S.; Crow, T.R.; Shifley, S.R. 2004. Fuel load reductions and fire risk in central hardwood forests of the United States: a spatial simulation study. Ecological Modelling. 180(1): 89-102. Sharitz, R.R.; Boring, L.R.; Lear, D.H.V.; Pinder, J.E., III. 1992. Integrating ecological concepts with natural resource management of southern forests. Ecological Applications. 2(3): 22-237. Shuman, B.; Bartlein, P.; Logar, N.; Newby, P.; Webb, T. 2002. Parallel climate and vegetation responses to the early Holocene collapse of the Laurentide ice sheet. Quaternary Science Reviews. 21(1): 1793-1805. Shumway, D.L.; Abrams, M.D.; Ruffner, C.M. 2001. A 400-year history of fire and oak recruitment in an old-growth oak forest in western Maryland, U.S.A. Canadian Journal of Forest Research. 31(8): 1437-1443. Simmons, J.A.; Currie, W.S.; Eshleman, K.N.; Kuers, K.; Monteleone, S.; Negley, T.L.; Pohlad, B.R.; Thomas, C.L. 2008. Forest to reclaimed mine land use change leads to altered ecosystem structure and function. Ecological Applications. 18(1): 104-118. Sinha, T.; Cherkauer, K.A. 2010. Impacts of future climate change on soil frost in the midwestern United States. Journal of Geophysical Research: Atmospheres (1984-2012). 115(D8). 22 Society of American Foresters. 2014. The dictionary of forestry. Bethesda, MD: Society of American Foresters. Available at http:// dictionaryofforestry.org/dict/term/ecosystem. (Accessed January 10, 2014). Soil Conservation Service [SCS]. 1974. Soil survey of Garrett County, Maryland. Washington, DC: U.S. Department of Agriculture, Soil Conservation Service and Maryland Agriculture Experiment Station. 87 p. Sommers, W.T.; Coloff, S.G.; Conard, S.G. 2011. Fire history and climate change. Report submitted to the Joint Fire Science Program for Project 09-02-1-09. 190 p. Available at https://www.firescience.gov/projects/09-2-019/project/09-2-01-9_09_2_01_9_Deliverable_ 01.pdf. (Accessed August 21, 2014). Souther, S.; McGraw, J.B. 2011. Evidence of local adaptation in the demographic response of American ginseng to interannual temperature variation. Conservation Biology. 25(5): 922-931. Souza, L.; Belote, R.; Kardol, P.; Weltzin, J.; Norby, R. 2010. CO2 enrichment accelerates successional development of an understory plant community. Journal of Plant Ecology. 3(1): 33-39. Spence, B.; Kunkel, C.; Schewel, E.; Boettner, T.; Martin, L. 2013. Who owns West Virginia? Charleston, WV: West Virginia Center on Budget and Policy and the American Friends Service Committee. 57 p. Available at http://www. wvpolicy.org/wp-content/uploads/2013/12/landstudy-paper-final3.pdf. (Accessed August 21, 2014). LiTERATuRE CiTED Stoner, A.M.K.; Hayhoe, K.; Yang, X.; Wuebbles, D.J. 2013. An asynchronous regional regression model for statistical downscaling of daily climate variables. International Journal of Climatology. 33(11): 2473-2494. Szlavecz, K.; McCormick, M.; Xia, L.; Saunders, J.; Morcol, T.; Whigham, D.; Filley, T.; Csuzdi, C. 2010. Ecosystem effects of non-native earthworms in Mid-Atlantic deciduous forests. Biological Invasions. 13(5): 115-1182. Stott, P.A.; Gillett, N.P.; Hegerl, G.C.; Karoly, D.J.; Stone, D.A.; Zhang, X.; Zwiers, F. 2010. Detection and attribution of climate change: a regional perspective. Wiley Interdisciplinary Reviews: Climate Change. 1(2): 192-211. Tans, P.; Keeling, R. 2013. Trends in atmospheric carbon dioxide. National Oceanic and Atmospheric Administration, Earth System Research Laboratory. Available at http://www.esrl.noaa.gov/gmd/ccgg/trends/. (Accessed August 21, 2014). Sturdevant-Rees, P.; Smith, J.A.; Morrison, J.; Baeck, M.L. 2001. Tropical storms and the flood hydrology of the Central Appalachians. Water Resources Research. 37(8): 2143-218. Sustainable Forestry Initiative [SFI]. 2013. SFI in Brief: April 2013 newsletter. Available at http://www.sfiprogram.org/media-resources/ newsletter1/april-2013/. (Accessed August 21, 2014). Swanson, F.J.; Johnson, S.L.; Gregory, S.V.; Acker, S.A. 1998. Flood disturbance in a forested mountain landscape. BioScience. 48(9): 81-89. Swanston, C.; Janowiak, M.; Iverson, L.; Parker, L.; Mladenoff, D.; Brandt, L.; Butler, P.; Pierre, M.S.; Prasad, A.; Matthews, S.; Peters, M.; Higgins, D.; Dorland, A. 2011. Ecosystem vulnerability assessment and synthesis: a report from the Climate Change Response Framework project in northern Wisconsin. Gen. Tech. Rep. NRS-82. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 142 p. Swanston, C.W.; Janowiak, M.K. 2012. Forest adaptation resources: climate change tools and approaches for land managers. Gen. Tech. Rep. NRS-87. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 121 p. Available at http://www.nrs.fs.fed. us/pubs/40543. (Accessed August 21, 2014). Taylor, D.A. 200. Ginseng, the divine root: the curious history of the plant that captivated the world. Chapel Hill, NC: Algonquin Books. 304 p. Tebaldi, C.; Adams-Smith, D.; Heller, N. 2012. The heat is on: US temperature trends. Princeton, NJ: Climate Central. 22 p. Available at http://www.climatecentral.org/wgts/heat-ison/HeatIsOnReport.pdf. (Accessed August 21, 2014). Thomas, R.B.; Spal, S.E.; Smith, K.R.; Nippert, J.B. 2013. Evidence of recovery of Juniperus virginiana trees from sulfur pollution after the Clean Air act. Proceedings of the National Academy of Sciences. 110(38): 15319-15324. Thomas-Van Gundy, M.A.; Nowacki, G.J. 2013. The use of witness trees as pyro-indicators for mapping past fire conditions. Forest Ecology and Management. 304(15): 333-344. Thomas-Van Gundy, M.A.; Strager, M.P. 2012. European settlement-era vegetation of the Monongahela National Forest, West Virginia. Gen. Tech. Rep. NRS-101. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 39 p. Tilman, D. 199. Biodiversity: population versus ecosystem stability. Ecology. 77(2): 350-33. 227 LiTERATuRE CiTED Tilman, D. 1999. The ecological consequences of changes in biodiversity: a search for general principles. Ecology. 80(5): 1455-1474. Tomer, M.D.; Schilling, K.E. 2009. A simple approach to distinguish land-use and climatechange effects on watershed hydrology. Journal of Hydrology. 37(1): 24-33. Townsend, P.A.; Helmers, D.P.; Kingdon, C.C.; McNeil, B.E.; de Beurs, K.M.; Eshleman, K.N. 2009. Changes in the extent of surface mining and reclamation in the Central Appalachians detected using a 1976-2006 Landsat time series. Remote Sensing of Environment. 113(1): 2-72. Townsend Peterson, A.; Scachetti-Pereira, R. 2004. Potential geographic distribution of Anoplophora glabripennis (Coleoptera: Cerambycidae) in North America. The American Midland Naturalist. 151(1): 170-178. Trapp, R.J.; Diffenbaugh, N.S.; Brooks, H.E.; Baldwin, M.E.; Robinson, E.D.; Pal, J.S. 2007. Changes in severe thunderstorm environment frequency during the 21st century caused by anthropogenically enhanced global radiative forcing. Proceedings of the National Academy of Sciences of the United States of America. 104(50): 19719-19723. Trapp, R.J.; Diffenbaugh, N.S.; Gluhovsky, A. 2009. Transient response of severe thunderstorm forcing to elevated greenhouse gas concentrations. Geophysical Research Letters. 3(1): L01703. Trapp, R.J.; Robinson, E.D.; Baldwin, M.E.; Diffenbaugh, N.S.; Schwedler, B.R. 2011. Regional climate of hazardous convective weather through high-resolution dynamical downscaling. Climate Dynamics. 37(3-4): 77-88. 228 Trout Unlimited. 200. Eastern Brook trout: status and threats. Arlington, VA. 40 p. Available at http://easternbrooktrout.org/reports/easternbrook-trout-status-and-threats/at_download/file. (Accessed August 21, 2014). Tu, M.; Callie, H.; M., R.J. 2001. Weed control methods handbook: tools and techniques for use in natural areas. Davis, CA: The Nature Conservancy. Available at http://www.invasive. org/gist/handbook.html. (Accessed August 21, 2014). Turcotte, R.M.; Elliott, T.R.; Fajvan, M.A.; Park, Y.-L.; Snider, D.A.; Tobin, P.C. 2012. Effects of ice storm damage on hardwood survival and growth in Ohio. Northern Journal of Applied Forestry. 29(2): 53-59. Ulbrich, U.; Leckebusch, G.; Pinto, J. 2009. Extratropical cyclones in the present and future climate: a review. Theoretical and Applied Climatology. 9(1-2): 117-131. Ulbrich, U.; Pinto, J.; Kupfer, H.; Leckebusch, G.; Spangehl, T.; Reyers, M. 2008. Changing Northern Hemisphere storm tracks in an ensemble of IPCC climate change simulations. Journal of Climate. 21(8): 19-179. Ungerer, M.J.; Ayres, M.P.; Lombardero, M.J. 1999. Climate and the northern distribution limits of Dendroctonus frontalis Zimmermann (Coleoptera: Scolytidae). Journal of Biogeography. 2(): 1133-1145. U.S. Department of Agriculture. 2012. USDA plant hardiness zone map. Agricultural Research Service. Available at http://planthardiness.ars. usda.gov/PHZMWeb/. (Accessed September 9, 2014). U.S. Department of Energy. 2012. A review of power outages and restoration following the June 2012 derecho. Infrastructure Security and Energy Restoration, Office of Electricity Delivery and Energy Reliability. Available at http:// energy.gov/sites/prod/files/Derecho%202012_ %20Review_0.pdf. (Accessed August 21, 2014). LiTERATuRE CiTED U.S. Environmental Protection Agency [EPA]. 1998. National water quality inventory: 1998 report to Congress 305b. Available at http://water.epa. gov/lawsregs/guidance/cwa/305b/98report_index. cfm. (Accessed 15 January 2014). U.S. Environmental Protection Agency [EPA]. 2005. Mountaintop mining/valley fills in Appalachia. Final Programmatic Environmental Impact Statement. Philadelphia, PA: EPA Region 3. 507 p. Available at http://nepis.epa.gov/Exe/ ZyPDF.cgi/20005XA.PDF?Dockey=20005XA. PDF. (Accessed August 21, 2014). U.S. Environmental Protection Agency [EPA]. 2008. National water program strategy: response to climate change. U.S. Environmental Protection Agency, Office of Water. Available at http://water. epa.gov/scitech/climatechange/upload/2008National-Water-Program-Strategy-Responseto-Climate-Change.pdf. (Accessed August 21, 2014). U.S. Environmental Protection Agency [EPA]. 2011. The effects of mountaintop mines and valley fills on aquatic ecosystems of the central Appalachian coalfields. EPA/00/R09/138F. Washington, DC: National Center for Environmental Assessment. Available at file:/// C:/Users/prbutler/Downloads/MOUNTAINTOP_ MINING_&_VALLEY_FILL_-_FINAL.PDF. (Accessed August 21, 2014). U.S. Forest Service [USFS]. 200a. Monongahela National Forest final environmental impact statement for plan revision. Elkins, WV: U.S. Department of Agriculture, Forest Service. 259 p. Available at http://www.fs.usda.gov/ Internet/FSE_DOCUMENTS/fsm9_01135.pdf. (Accessed August 21, 2014). U.S. Forest Service. 200b. Wayne National Forest 2006 forest plan. Land and Resource Management Plan (200 Forest Plan). Milwaukee, WI: U.S. Department of Agriculture, Forest Service, Eastern Region. 30 p. Available at http://www.fs.usda.gov/Internet/FSE_ DOCUMENTS/fsm9_00523.pdf. (Accessed August 21, 2014). U.S. Forest Service. 2007. USDA Forest Service manual 2300: recreation, wilderness, and related resource management. Chapter 2320: Wilderness Management. Washington, DC: U.S. Department of Agriculture, Forest Service, National Headquarters (WO). 5 p. Available at http://www.fs.usda.gov/Internet/FSE_ DOCUMENTS/fsbdev3_053277.pdf. (Accessed August 21, 2014). U.S. Forest Service. 2008. Who owns America’s forests? 8 p. Available at http://www.nrs.fs.fed. us/pubs/inf/NRS-INF-0-08.pdf. (Accessed August 21, 2014). U.S. Fish and Wildlife Service [USFWS]. 2010. Turkey hunting in 2006: an analysis of hunter demographics, trends, and economic impacts. Arlington, VA. 2 p. U.S. Forest Service. 2011a. National report on sustainable forests - 2010. FS 979. Available at http://www.fs.fed.us/research/sustain/docs/ national-reports/2010/2010-sustainability-report. pdf. (Accessed August 21, 2014). U.S. Fish and Wildlife Service [USFWS]. 2014. Engangered species database. Available at http://www.fws.gov/endangered/. (Accessed May 14, 2014). U.S. Forest Service. 2011b. National visitor use monitoring results. Available at http://apps.fs.fed.us/nrm/nvum/results/. (Accessed August 22, 2014). 229 LiTERATuRE CiTED U.S. Forest Service. 2013. FIA DataMart: FIADB version 5.1. Microsoft Access 2007. Online database. Washington, DC: U.S. Department of Agriculture, Forest Service, Forest Inventory and Analysis Program. Available at http://apps.fs.fed. us/fiadb-downloads/datamart.html. (Accessed April 11, 2014). U.S. Geological Survey [USGS]. 1989. West Virginia floods and droughts. Available at http://md.water.usgs.gov/publications/wsp-2375/ wv/. (Accessed September 9, 2014). U.S. Geological Survey [USGS]. 2011. National Land Cover Dataset 2006 land cover. Available at http://www.mrlc.gov/index.php. (Accessed December 8, 2011). Vadas, R.L., Jr.; Sanger, J.E. 1997. Lateral zonation of trees along a small Ohio stream. Ohio Journal of Science. 97(5): 107-112. van der Werf, G.R.; Morton, D.C.; DeFries, R.S.; Olivier, J.G.; Kasibhatla, P.S.; Jackson, R.B.; Collatz, G.J.; Randerson, J. 2009. CO2 emissions from forest loss. Nature Geoscience. 2(11): 737-738. Vanderhorst, J.; Streets, B.; Jeuck, J.; Gawler, S. 2008. Vegetation classification and mapping of Bluestone National Scenic River, West Virginia. Technical Report NPS/NER/NRTR2008/10. Philadelphia, PA: U.S. Department of the Interior, National Park Service. Available at http://irmafiles.nps.gov/reference/holding/ 150158?accessType=DOWNLOAD. (Accessed August 21, 2014). Venette, R.C.; Abrahamson, M. 2010. Cold hardiness of emerald ash borer, Agrilus planipennis: a new perspective. Black ash symposium: proceedings of the meeting; May 25-27, 2010; Bemidji, MN. Cass Lake, MN: U.S. Department of Agriculture, Forest Service, Chippewa National Forest. 5 p. Available at http://www.treesearch.fs.fed.us/pubs/45307. (Accessed September 18, 2014). 230 Villarini, G.; Smith, J.A.; Baeck, M.L.; Krajewski, W.F. 2011. Examining flood frequency distributions in the Midwest U.S. Journal of the American Water Resources Association. 47(3): 447-43. Vitousek, P.M.; Porder, S.; Houlton, B.Z.; Chadwick, O.A. 2010. Terrestrial phosphorus limitation: mechanisms, implications, and nitrogen-phosphorus interactions. Ecological Applications. 20(1): 5-15. Vose, J.; Clinton, B.; Swank, W. 1993. Fire, drought, and forest management influences on pine/hardwood ecosystems in the southern Appalachians. A paper presented at the 12th conference on fire and forest meteorology; 1993 October 2-28; Jekyll Island, GA. 232-238. Available at http://www.treesearch.fs.fed.us/ pubs/472. (Accessed August 21, 2014). Vose, J.M.; Peterson, D.L.; Patel-Weynand, T. 2012. Effects of climatic variability and change on forest ecosystems: a comprehensive science synthesis for the US forest sector. Gen. Tech. Rep. PNW-GTR-870. U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 282 p. Wagner, R.J.; Kaye, M.W.; Abrams, M.D.; Hanson, P.J.; Martin, M. 2012. Tree-ring growth and wood chemistry response to manipulated precipitation variation for two temperate Quercus species. Tree-Ring Research. 8(1): 17-29. Walck, J.L.; Hidayati, S.N.; Dixon, K.W.; Thompson, K.E.N.; Poschlod, P. 2011. Climate change and plant regeneration from seed. Global Change Biology. 17(): 2145-211. Walker, B.; Kinzig, A.; Langridge, J. 1999. Original articles: plant attribute diversity, resilience, and ecosystem function: the nature and significance of dominant and minor species. Ecosystems. 2(2): 95-113. LiTERATuRE CiTED Walker, B.H. 1992. Biodiversity and ecological redundancy. Conservation Biology. (1): 18-23. Waller, D.M.; Alverson, W.S. 1997. The whitetailed deer: a keystone herbivore. Wildlife Society Bulletin. 25(2): 217-22. Walters, B.F.; Rider, D.R.; Piva, R.J. 2008. Maryland timber industry: an assessment of timber product output and use, 2008. Resour. Bull. NRS-4. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 52 p. Ward, N.L.; Masters, G.J. 2007. Linking climate change and species invasion: an illustration using insect herbivores. Global Change Biology. 13(8): 105-115. Washington, W.M.; Weatherly, J.W.; Meehl, G.A.; Semtner, A.J., Jr.; Bettge, T.W.; Craig, A.P.; Strand, W.G., Jr.; Arblaster, J.; Wayland, V.B.; James, R.; Zhang, Y. 2000. Parallel climate model (PCM) control and transient simulations. Climate Dynamics. 1(10): 755-774. Walther, G.R.; Post, E.; Convey, P.; Menzel, A.; Parmesan, C.; Beebee, T.J.C.; Fromentin, J.M.; Hoegh-Guldberg, O.; Bairlein, F. 2002. Ecological responses to recent climate change. Nature. 41(879): 389-395. Weary, D.J. 2008. Preliminary map of potentially karstic carbonate rocks in the Central and Southern Appalachian states [Online]. Open File Report 2008-1154. U.S. Geological Survey. Available at http://pubs.usgs.gov/of/2008/1154/. (Accessed August 21, 2014). Wang, D.; Cai, X. 2010. Comparative study of climate and human impacts on seasonal baseflow in urban and agricultural watersheds. Geophysical Research Letters. 37(). Webb, T.P.; Bartlein, J. 1992. Global changes during the last 3 million years: climatic controls and biotic responses. Annual Review of Ecology and Systematics. 23(1): 141. Wang, J.; Bai, X.; Hu, H.; Clites, A.; Colton, M.; Lofgren, B. 2012. Temporal and spatial variability of Great Lakes ice cover, 1973-2010. Journal of Climate. 25(4): 1318-1329. Weed, A.S.; Ayres, M.P.; Hicke, J. 2013. Consequences of climate change for biotic disturbances in North American forests. Ecological Monographs. 83: 441-470. Wang, J.; Zhang, X. 2008. Downscaling and projection of winter extreme daily precipitation over North America. Journal of Climate. 21(5): 923-937. Wei, X.; Wu, Y.; Reardon, R.; Sun, T.H.; Lu, M.; Sun, J.H. 2007. Biology and damage traits of emerald ash borer (Agrilus planipennis Fairmaire) in China. Insect Science. 14(5): 37-373. Wang, W.; He, H.S.; Spetich, M.A.; Shifley, S.R.; Thompson, F.R.; Larsen, D.R.; Fraser, J.S.; Yang, J. 2013. A large-scale forest landscape model incorporating multi-scale processes and utilizing forest inventory data. Ecosphere. 4(9): art10. West Virginia Department of Environmental Protection [WVDEP]. 2013a. Office of Oil and Gas [Web site]. Available at http://www.dep. wv.gov/oil-and-gas/Pages/default.aspx. (Accessed March 7, 2013). Wang, W.J.; He, H.S.; Fraser, J.S.; Thompson, F.R.; Shifley, S.R.; Spetich, M.A. 2014. LANDIS PRO: a landscape model that predicts forest composition and structure changes at regional scales. Ecography. 37: 225-229. West Virginia Department of Environmental Protection [WVDEP]. 2013b. West Virginia water resources management plan. Charleston, WV. Available at http://www.dep.wv.gov/WWE/ wateruse/WVWaterPlan/Documents/WV_ WRMP.pdf. (Accessed August 21, 2014). 231 LiTERATuRE CiTED West Virginia Division of Forestry. 2010. West Virginia statewide forest resource assessment. Charleston, WV. 57 p. West Virginia Division of Forestry. 2012. Managed timberland fact sheet. 2 p. Available at http:// www.wvforestry.com/MT%20Fact%20Sheet.pdf. (Accessed August 21, 2014). West Virginia Division of Natural Resources [WVDNR]. 2014. Rare, threatened, and endangered species in West Virginia. Wildlife Resources Section. Available at http://www. wvdnr.gov/Wildlife/PDFFiles/RTE_Animals_ 2012.pdf. (Accessed January 14, 2014). Wickham, J.; Wood, P.B.; Nicholson, M.C.; Jenkins, W.; Druckenbrod, D.; Suter, G.W.; Strager, M.P.; Mazzarella, C.; Galloway, W.; Amos, J. 2013. The overlooked terrestrial impacts of mountaintop mining. BioScience. 3(5): 335-348. Wiedenbeck, J.K.; Sabula, A. 2008. Ohio roundwood utilization by the timber industry in 2006. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 18 p. Wiens, J.A.; Stralberg, D.; Jongsomjit, D.; Howell, C.A.; Snyder, M.A. 2009. Niches, models, and climate change: assessing the assumptions and uncertainties. Proceedings of the National Academy of Sciences. 10(Supplement 2): 19729-1973. Wilderness Act of 1964. 1 U.S.C. 1121 (note), 1131-113. Available at http://www.fs.usda.gov/ Internet/FSE_DOCUMENTS/stelprdb5253940. pdf. (Accessed August 21, 2014). Wiley, J.B.; Atkins, J.T. 2010. Estimation of floodfrequency discharges for rural, unregulated streams in West Virginia. Scientific Investigations Report 2010-5033. Reston, VA: U.S. Department of the Interior, U.S. Geological Survey. 88 p. Widmann, R.H.; Balser, D.; Barnett, C.J.; Butler, B.J.; Griffith, D.M.; Lister, T.W.; Moser, W.K.; Perry, C.H.; Riemann, R.; Woodall, C. 2009. Ohio forests: 2006. Resour. Bull. NRS-3. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 119 p. Williams, C.E. 2005. Fact sheet: exotic bush honeysuckles. Clarion, PA: Plant Conservation Alliance’s Alien Plant Working Group. Available at http://www.nps.gov/plants/alien/fact/pdf/loni1. pdf. (Accessed August 21, 2014). Widmann, R.H.; Dye, C.R.; Cook, G.W. 2007. Forests of the Mountain State. Resour. Bull. NRS-17. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 32 p. Willis, C.G.; Ruhfel, B.R.; Primack, R.B.; MillerRushing, A.J.; Losos, J.B.; Davis, C.C. 2010. Favorable climate change response explains non-native species’ success in Thoreau’s woods. PLoS ONE. 5(1): e8878. Widmann, R.H.; Morin, R.S. 2012. Ohio’s forest resources, 2012. Res. Note NRS-187. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 4 p. Available at http://www.fs.fed.us/nrs/pubs/rn/rn_ nrs187.pdf. (Accessed August 21, 2014). Winkler, J.A.; Arritt, R.; Pryor, S. 2012. Climate projections for the Midwest: availability, interpretation and synthesis. In: Winkler, J.; Andresen, J.; Hatfield, J.; Bidwell, D.; Brown, D., eds. U.S. National Climate Assessment Midwest technical input report. 24 p. Available at http:// glisa.umich.edu/docs/NCA/MTIT_Future.pdf. (Accessed August 21, 2014). 232 LiTERATuRE CiTED Wisconsin Initiative on Climate Change Impacts [WICCI]. 2011. Wisconsin’s changing climate: impacts and adaptation. Madison, WI. 224 p. Wright, H.A.; Bailey, A.W. 1982. Fire ecology: United States and southern Canada. New York, NY: John Wiley & Sons. 528 p. Woodall, C.W.; Oswalt, C.M.; Westfall, J.A.; Perry, C.H.; Nelson, M.D.; Finley, A.O. 2009. An indicator of tree migration in forests of the eastern United States. Forest Ecology and Management. 257(5): 1434-1444. Wu, S.; Mickley, L.J.; Leibensperger, E.M.; Jacob, D.J.; Rind, D.; Streets, D.G. 2008. Effects of 2000-2050 global change on ozone air quality in the United States. Journal of Geophysical Research. 113(D0302): 1-12. Woods, A.J.; Omernik, J.M.; Brown, D.D. 1999. Level III and IV ecoregions of Delaware, Maryland, Pennsylvania, Virginia, and West Virginia. Corvallis, OR: U.S. Environmental Protection Agency, National Health and Environmental Effects Research Laboratory. Available at http://www.epa.gov/wed/pages/ ecoregions/reg3_eco.htm. (Accessed August 21, 2014). Wuebbles, D.J.; Hayhoe, K. 2004. Climate change projections for the United States Midwest. Mitigation and Adaptation Strategies for Global Change. 9: 335-33. Woods, A.J.; Omernik, J.M.; Brown, D.D.; Kiilsgaard, C.W. 199. Level III and IV ecoregions of Pennsylvania and the Blue Ridge Mountains, the Ridge and Valley, and Central Appalachians of Virginia, West Virginia, and Maryland. EPA/00/R-0/077. Washington, DC: U.S. Environmental Protection Agency, National Health and Environmental Effects Research Laboratory. 50 p. World Meteorological Organization. 2011. WMO statement on the status of the global climate in 2011. Geneva, Switzerland. Available at www.wmo.int/pages/prog/wcp/wcdmp/ documents/1085_en.pdf. (Accessed May 10, 2013). Wright, D.M.; Posselt, D.J.; Steiner, A.L. 2013. Sensitivity of lake-effect snowfall to lake ice cover and temperature in the Great Lakes region. Monthly Weather Review. 141(2): 70-89. Wullschleger, S.D.; Gunderson, C.A.; Tharp, M.L.; West, D.C.; Post, W.M. 2003. Simulated patterns of forest succession and productivity as a consequence of altered precipitation. In: Hanson, P.J.; Wullschleger, S.D., eds. North American temperate deciduous forest responses to changing precipitation regimes. New York, NY: Springer: 433-44. Whitney, G.G. 1994. From coastal wilderness to fruited plain: a history of environmental change in temperate North America 1500 to the present. New York, NY: Cambridge University Press. 451 p. Zhang, D.; Butler, B.J.; Nagubadi, R.V. 2012. Institutional timberland ownership in the US South: magnitude, location, dynamics, and management. Journal of Forestry. 110(7): 355-31. Zhang, X.; Tarpley, D.; Sullivan, J.T. 2007. Diverse responses of vegetation phenology to a warming climate. Geophysical Research Letters. 34(19): L19405. Ziska, L.H. 2003. Evaluation of the growth response of six invasive species to past, present and future atmospheric carbon dioxide. Journal of Experimental Botany. 54(381): 395-404. 233 APPENDiX 1. SPECiES LiSTS Table 24.—Common and scieniic names of naive plants menioned in this assessment Common Name Scieniic Name Common Name Scieniic Name balsam ir Abies balsamea silky dogwood Cornus amomum boxelder Acer negundo roughleaf dogwood Cornus drummondii black maple Acer nigrum lowering dogwood Cornus lorida striped maple Acer pensylvanicum ram’s-head lady’s-slipper Cypripedium arieinum red maple Acer rubrum common persimmon Diospyros virginiana silver maple Acer saccharinum American beech Fagus grandifolia sugar maple Acer saccharum white ash Fraxinus americana mountain maple Acer spicatum black ash Fraxinus nigra northern wild monkshood Aconitum noveboracense green ash Fraxinus pennsylvanica yellow buckeye Aesculus lava blue ash Fraxinus quadrangulata Ohio buckeye Aesculus glabra water locust Gleditsia aquaica lillydale onion Allium oxyphilum honeylocust Gleditsia triacanthos speckled alder Alnus incana bushy St. Johnswort Hypericum densilorum hazel alder Alnus serrulata American holly Ilex opaca serviceberry Amelanchier Medik. common winterberry Ilex vericillata shale barren rockcress Arabis seroina small whorled pogonia Isotria medeoloides pawpaw Asimina triloba buternut Juglans cinerea yellow birch Betula alleghaniensis black walnut Juglans nigra sweet birch Betula lenta eastern redcedar Juniperus virginiana river birch Betula nigra mountain laurel Kalmia laifolia American hornbeam Carpinus caroliniana tamarack Larix laricina mockernut hickory Carya alba northern spicebush Lindera benzoin biternut hickory Carya cordiformis sweetgum Liquidambar styracilua pignut hickory Carya glabra tulip tree Liriodendron tulipifera shagbark hickory Carya ovata osage orange Maclura pomifera black hickory Carya texana cucumbertree Magnolia acuminata American chestnut Castanea dentata mountain magnolia Magnolia fraseri northern catalpa Catalpa speciosa southern magnolia Magnolia grandilora sugarberry Celis laevigata red mulberry Morus rubra common hackberry Celis occidentalis blackgum Nyssa sylvaica common butonbush Cephalanthus occidentalis eastern hophornbeam Ostrya virginiana eastern redbud Cercis canadensis sourwood Oxydendrum arboreum Bentley’s coralroot Corallorhiza bentleyi American ginseng Panax quinquefolius 234 APPENDiX 1 Table 24 (coninued). Common Name Scieniic Name Common Name Scieniic Name black chokeberry Phoinia melanocarpa chinkapin oak Quercus muehlenbergii red spruce Picea rubens water oak Quercus nigra shortleaf pine Pinus echinata pin oak Quercus palustris Table Mountain pine Pinus pungens willow oak Quercus phellos red pine Pinus resinosa chestnut oak Quercus prinus pitch pine Pinus rigida northern red oak Quercus rubra eastern white pine Pinus strobus Shumard’s oak Quercus shumardii loblolly pine Pinus taeda post oak Quercus stellata Virginia pine Pinus virginiana black oak Quercus veluina eastern prairie fringed orchid Platanthera leucophaea great laurel Rhododendron maximum black locust Robinia pseudoacacia sycamore Platanus occidentalis coastal plain willow Salix caroliniana eastern cotonwood Populus deltoides black willow Salix nigra bigtooth aspen Populus grandidentata Sassafras albidum quaking aspen sassafras Populus tremuloides northeastern bulrush Scirpus ancistrochaetus Tennessee pondweed Potamogeton tennesseensis American mountain ash Sorbus americana pin cherry Prunus pensylvanica Prunus seroina Virginia spiraea Spiraea virginiana black cherry Thuja occidentalis chokecherry northern white-cedar Prunus virginiana Tilia americana Pilimnium nodosum American basswood harperella running bufalo clover Trifolium stoloniferum Torrey’s mountainmint Pycnanthemum torrei eastern hemlock Tsuga canadensis white oak Quercus alba winged elm Ulmus alata swamp white oak Quercus bicolor American elm Ulmus americana scarlet oak Quercus coccinea cedar elm Ulmus crassifolia northern pin oak Quercus ellipsoidalis slippery elm Ulmus rubra southern red oak Quercus falcata rock elm Ulmus thomasii bear oak/scrub oak Quercus ilicifolia Vaccinium erythrocarpum shingle oak Quercus imbricaria southern mountain cranberry bur oak Quercus macrocarpa velvetleaf huckleberry Vaccinium myrilloides blackjack oak Quercus marilandica wild raisin (withe-rod) Viburnum nudum 235 APPENDiX 1 Table 25.—Common and scieniic names of pathogens and nonnaive plants menioned in this assessment Common Name Scieniic Name Pathogens Common Name Scieniic Name Pathogens armillaria Armillaria mellea scleroderris canker Gremmeniella abieina Lyme disease Borrelia burgdorferi annosum root disease Heterobasidion irregulare elm yellows Candidatus phytoplasma ulmi hypoxylon canker Hypoxylon mammatum white pine blister rust Cronarium ribicola sudden oak death Phytophthora ramorum chestnut blight Cryphonectria parasiica phytophthora root rot Phytophthora spp. diplodia Diplodia pinea and D. scrobiculata sirococcus shoot blight Sirococcus conigenus sphaeropsis shoot blight Sphaeropsis sapinea West Nile virus Common Name Flavivirus spp. Scieniic Name Common Name Scieniic Name Nonnaive invasive plants Nonnaive invasive plants Norway maple Acer platanoides cogongrass Imperata cylindrica ailanthus Ailanthus alissima sericea lespedeza Lespedeza cuneata silk tree Albizia julibrissin privet Ligustrum vulgare garlic mustard Alliaria peiolata Japanese honeysuckle Lonicera japonica porcelain berry Ampelopsis brevipedunculata bush honeysuckle Lonicera mackii purple loosestrife Lythrum salicaria dwarf mistletoe Arceuthobium pusillum yellow sweetclover Melilotus oicinalis Japanese barberry Berberis thunbergii Broussoneia papyrifera Japanese siltgrass Microstegium vimineum paper mulberry basket grass Oplismenus hirtellus Asiaic bitersweet Celastrus orbiculatus princess tree Paulownia tomentusa spoted knapweed Centaurea stoebe Dennstaedia puncilobula mile-a-minute vine Persicaria perfoliata hayscented fern reed canarygrass Phalaris arundinacea viper’s bugloss Echium vulgare common reed (phragmites) Phragmites australis autumn olive Elaeagnus umbellata Canada bluegrass Poa compressa burning bush Euonymus spp. kudzu Pueraria lobata Japanese knotweed Fallopia japonica glossy buckthorn Rhamnus spp. buckthorn Frangula alnus mulilora rose Rosa mulilora creeping charlie Glechoma hederacea crown vetch Securigera varia English ivy Hedera helix Japanese spiraea Spiraea japonica 23 APPENDiX 1 Table 26.—Common and scieniic names of fauna menioned in this assessment Common Name Scieniic Name Common Name Scieniic Name hemlock woolly adelgid Adelges tsugae bark beetle saw-whet owl Aegolius acadicus Ips spp. and Dendroctonus spp. emerald ash borer Agrilus planipennis spring hemlock looper Jeferson salamander Ambystoma jefersonianum Lambdina iscellaria iscellaria spoted salamander Ambystoma maculata silver-haired bat Lasionycteris nocivagans iger salamander Ambystoma igrinum eastern red bat Lasiurus borealis green salamander Aneides aeneus hoary bat Lasiurus cinereus Asian longhorned beetle Anoplophora glabripennis crimson-ringed whiteface Leucorrhinia glacialis Dry Fork Valley cave pseudoscorpion Apochthonius pauscisinosus red crossbill Loxia curvirostra gypsy moth Lymantria dispar dispar eastern cave-loving funnel web spider Calymmaria cavicola bobcat Lynx rufus forest tent caterpillar Malacosoma disstria coyote Canis latrans wild turkey Meleagris gallopavo eastern imber wolf Canis lupus lycaon small-footed bat Myois leibii beaver Castor canadensis litle brown bat Myois lucifugus spruce budworm Choristoneura fumiferana northern bat Myois septentrionalis spoted turtle Clemmys gutata Indiana bat Myois sodalis redside dace Clinostomus elongatus Carter cave spider Nesicus carteri Virginia big-eared bat Corynorhinus townsendii virginianus jumping oak gall wasp Neuroterus sp. Notophthalmus viridescens Cotus spp. red-spoted newt sculpin white-tailed deer Odocoileus virginianus eastern hellbender Cryptobranchus alleganiensis Kentucky warbler Oporornis formosus beech scale Cryptococcus fagisuga tri-colored bat Perimyois sublavus earthworms (nonnaive) Dendrobaena octaedra, Lumbricus rubellus, and L. terrestris red-backed salamander Plethodon cinereus Cheat Mountain salamander Plethodon neingi southern pine beetle Dendroctonus frontalis eastern cougar Puma concolor couguar blackburnian warbler Dendroica fusca northern lying squirrel Sabrinus glaucomys fuscus birch leaf miner Fenusa pusilla brook trout Salvelinus foninalis southern lying squirrel Glaucomys volans eastern spadefoot toad Scaphiopus holbrookii bog turtle Glyptemys muhlenbergii eastern gray squirrel Sciurus carolinensis rapids clubtail Gomphus quadricolor cerulean warbler Setophaga cerulea green-faced clubtail Gomphus viridifrons black bear Ursus americanus worm-eaing warbler Helmitheros vermivorus ambrosia beetle Xyloterinus politus wood thrush Hylocichla mustelina 237 APPENDiX 2: TREND ANALYSiS AND hiSToRiCAL CLiMATE DATA We used the ClimateWizard Custom Analysis tool to examine historical averages and trends in precipitation and temperature within the assessment area (Gibson et al. 2002, Girvetz et al. 2009). Data for ClimateWizard are derived from PRISM (Parameter-elevation Regressions on Independent Slopes Model) (Gibson et al. 2002). The PRISM model interpolates historical data from the National Weather Service cooperative stations, the Midwest Climate Data Center, and the Historical Climate Network, among others. Data undergo strict quality control procedures to check for errors in station measurements. The PRISM model finds linear relationships between these station measurements and local elevation by using a digital elevation model (digital gridded version of a topographic map). Temperature and precipitation are then derived for each pixel on a continuous 2.5-mile grid across the conterminous United States. The closer a station is to a grid cell of interest in distance and elevation, and the more similar it is in its proximity to coasts or topographic features, the higher the weight the station will have on the final, predicted value for that cell. More information on PRISM can be found at: www.prism.oregonstate.edu/. Please note that Web addresses are current as of the publication date of this assessment but are subject to change. A 30-year climate “normal” for the assessment area and each ecological section within the assessment area was calculated from the mean for 1971 through 2000 (Table 27). Linear trend analysis was 238 performed for 1901 through 2011 by using restricted maximum likelihood (REML) estimation (Girvetz et al. 2009). Restricted maximum likelihood methods were used for trend analysis of past climate for the International Panel on Climate Change Working Group 1 Report and are considered an effective way to determine trends in climate data over time (Trenberth et al. 2007). A first-order autoregression was assumed for the residuals, meaning that values one time step away from each other are assumed to be correlated. This method was used to examine trends for every 2.5-mile grid cell. The slope and p-values for the linear trend over time were calculated annually, seasonally, and monthly for each climate variable, and then mapped. An overall trend for an area is based on the trend analysis of the average value for all grid cells within the area over time (Table 28). The developers of the ClimateWizard tool advise users to interpret the linear trend maps in relation to the respective map of statistical confidence (Figs. 44 and 45). In this case, statistical confidence is described by using p-values from a t-test applied to the linear regression. A p-value can be interpreted as the probability of the slope being different from zero by chance. For this assessment, p-values of less than 0.1 were considered to have sufficient statistical confidence. Areas with low statistical confidence in the rate of change (gray areas on the map) should be interpreted with caution. APPENDiX 2 Table 27.—Annual and seasonal mean values for selected climate variables from 1971 through 2000 for ecological secions within the assessment area Ecological secion Season Precipitaion (inches) Mean temperature (°F) Minimum temperature (°F) Maximum temperature (°F) 221E Annual Fall Spring Summer Winter 42.3 9.3 11.3 12.7 9.1 52.1 53.9 51.3 71.3 32.0 40.7 42.2 38.8 59.6 22.3 63.5 65.6 63.8 83.0 41.6 221F Annual Fall Spring Summer Winter 39.6 9.6 10.3 12.0 7.7 49.3 51.8 48.1 69.4 28.0 39.2 41.6 37.1 58.2 19.8 59.5 62.0 59.1 80.7 36.2 M221A Annual Fall Spring Summer Winter 39.1 9.6 10.5 11.2 7.8 51.3 53.0 50.3 70.1 31.9 39.8 41.1 38.0 58.0 21.9 62.9 64.9 62.6 82.2 41.9 M221B Annual Fall Spring Summer Winter 48.5 10.7 13.1 13.7 10.9 49.0 50.7 48.1 66.8 30.3 37.8 39.3 36.0 55.5 20.4 60.2 62.2 60.2 78.1 40.1 M221C Annual Fall Spring Summer Winter 47.0 10.0 12.7 13.7 10.5 52.4 53.9 51.9 70.2 33.8 41.0 42.3 39.1 58.9 23.8 63.9 65.5 64.6 81.6 43.8 In addition, because maps are developed from weather station observations that have been spatially interpolated, developers of the ClimateWizard tool and PRISM data set recommend that inferences about trends should not be made for single grid cells or even small clusters of grid cells. The number of weather stations has also changed over time, and station data are particularly limited before 1948, meaning grid cells from earlier in the century are based on an interpolation of fewer points than later in the century (Gibson et al. 2002). Therefore, interpretations should be based on many grid cells showing regional patterns of climate change with high statistical confidence. For those interested in understanding trends in climate at a particular location, it is best to refer to weather station data for the closest station in the Global Historical Climatology Network from the National Climatic Data Center (http://www.ncdc.noaa.gov/). We selected the time period 1901 through 2011 because it was sufficiently long to capture interdecadal and intradecadal variation in climate for the region. We acknowledge that different trends can be inferred by selecting different beginning and end points in the analysis. Therefore, trends should be interpreted based on their relative magnitude and direction, and the slope of any single trend should be interpreted with caution. 239 APPENDiX 2 Table 28.—Annual, seasonal, and monthly mean values and linear trend analysis for selected climate variables from 1901 through 2011 for the assessment area. Month or season Mean Precip. Mean precip. change Precip. TMean (inches) (inches) p-valuea (°F) TMean Mean change TMean TMin (°F) p-valuea (°F) TMin change TMin (°F) p-valuea Mean TMax (°F) TMax change TMax (°F) p-valuea January February March April May June July Aug Sept Oct Nov Dec 3.3 2.8 3.8 3.7 4.1 4.2 4.5 3.9 3.2 2.8 3.0 3.2 -0.7 -0.1 -0.4 0.2 0.9 -0.4 0.3 -0.2 0.9 0.3 1.2 -0.1 0.14 0.70 0.40 0.51 0.05 0.38 0.33 0.47 0.04 0.48 0.01 0.70 29.7 31.4 40.3 50.5 60.0 68.2 71.9 70.5 64.3 53.2 42.2 32.6 -2.4 1.7 -0.1 2.4 0.0 0.5 0.0 1.2 -0.6 -0.8 2.3 1.5 0.24 0.38 0.94 0.00 0.96 0.50 0.97 0.03 0.51 0.42 0.01 0.25 20.4 21.2 29.1 38.0 47.4 56.0 60.1 58.8 52.0 40.7 31.6 23.6 -1.6 2.0 -0.3 1.6 0.6 1.4 1.3 2.1 0.9 0.4 2.8 1.7 0.41 0.33 0.79 0.02 0.42 0.07 0.06 0.00 0.40 0.73 0.00 0.19 39.1 41.5 51.5 62.9 72.7 80.3 83.7 82.3 76.6 65.7 52.8 41.6 -3.1 1.4 0.1 3.2 -0.7 -0.4 -1.2 0.3 -2.1 -2.0 1.8 1.2 0.14 0.48 0.95 0.00 0.48 0.68 0.10 0.64 0.03 0.07 0.08 0.37 Winter Spring Summer Fall Annual 3.1 3.9 4.2 3.0 42.4 -1.0 0.7 -0.3 2.3 1.7 0.14 0.29 0.64 0.00 0.26 31.2 50.3 70.2 53.2 51.2 0.3 0.8 0.6 0.3 0.5 0.81 0.19 0.22 0.57 0.29 21.7 38.2 58.3 41.4 39.9 0.7 0.6 1.6 1.4 1.1 0.58 0.24 0.00 0.04 0.03 40.7 62.4 82.1 65.0 62.6 -0.1 0.9 -0.4 -0.7 -0.1 0.91 0.20 0.46 0.27 0.87 *P-values represent the probability of observing that trend by chance. P-values in boldface indicate a less than 10-percent probability that the trend was due to chance. TMean = mean temperature, TMin = minimum temperature, TMax = maximum temperature. REFERENCES ClimateWizard. 2013. Climate Wizard Custom Analysis application. The Nature Conservancy, University of Washington, University of Southern Mississippi, Climate Central, Santa Clara University. Available at http:// climatewizardcustom.org/. (Accessed May 2013). Gibson, W.P.; Daly, C.; Kittel, T.; Nychka, D.; Johns, C.; Rosenbloom, N.; McNab, A.; Taylor, G. 2002. Development of a 103-year high-resolution climate data set for the conterminous United States. Proceedings of the 13th American Meteorological Society Conference on Applied Climatology: 181-183. 240 Girvetz, E.H.; Zganjar, C.; Raber, G.T.; Maurer, E.P.; Kareiva, P.; Lawler, J.J. 2009. Applied climatechange analysis: the Climate Wizard tool. PLoS ONE. 4(12): e8320. Trenberth, K.E.; Jones, P.D.; Ambenje, P.; Bojariu, R.; Easterling, D.; Tank, A.K.; Parker, D.; Rahmizadeh, F.; Renwick, J.A.; Rusticucci, M.; Soden, B.; Zhai, P. 2007. Observations: surface and atmospheric climate change. In: Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K.B.; Tignor, M.; Miller, H.L., eds. Climate change 2007: the physical science basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York, NY: Cambridge University Press: 234-33. APPENDiX 2 Figure 44.—Staisical conidence (p-values for the linear regression) for trends in temperature from 1901 through 2011. Gray values represent areas of low staisical conidence. Data source: ClimateWizard (2013). 241 APPENDiX 2 Figure 45.—Staisical conidence (p-values for the linear regression) for trends in precipitaion from 1901 through 2011. Gray values represent areas of low staisical conidence. Data source: ClimateWizard (2013). 242 APPENDiX 3: ADDiTioNAL FuTuRE CLiMATE PRoJECTioNS This appendix provides supplementary information to Chapter 4, presented as maps of projected change for early- and mid-century (Figs. 4 through 53) and graphs of early-, mid-, and late-century departures from baseline climate (Figs. 54 through 58). Figure 46.—Projected diference in daily mean temperature at the beginning of the century (2010 through 2039) compared to baseline (1971 through 2000) for two climate scenarios. 243 APPENDiX 3 Figure 47.—Projected diference in daily minimum temperature at the beginning of the century (2010 through 2039) compared to baseline (1971 through 2000) for two climate scenarios. 244 APPENDiX 3 Figure 48.—Projected diference in daily maximum temperature at the beginning of the century (2010 through 2039) compared to baseline (1971 through 2000) for two climate scenarios. 245 APPENDiX 3 Figure 49.—Projected diference in precipitaion at the beginning of the century (2010 through 2039) compared to baseline (1971 through 2000) for two climate scenarios. 24 APPENDiX 3 Figure 50.—Projected diference in daily mean temperature for the middle of the century (2040 through 2069) compared to baseline (1971 through 2000) for two climate scenarios. 247 APPENDiX 3 Figure 51.—Projected diference in daily minimum temperature for the middle of the century (2040 through 2069) compared to baseline (1971 through 2000) for two climate scenarios. 248 APPENDiX 3 Figure 52.—Projected diference in daily maximum temperature for the middle of the century (2040 through 2069) compared to baseline (1971 through 2000) for two climate scenarios. 249 APPENDiX 3 Figure 53.—Projected diference in precipitaion for the middle of the century (2040 through 2069) compared to baseline (1971 through 2000) for two climate scenarios. 250 APPENDiX 3 W in te r (D e c - F e b ) 50 T e m p e ra tu re (0 F ) 45 40 35 30 25 20 1971 - 2000 2010 - 2039 2040 - 2069 2070 - 2099 GFD L A 1FI Mean G F D L A 1 F I M a xim u m G F D L A 1 F I M in im u m P C M B 1 Mean P C M B 1 M a xim u m P C M B 1 M in im u m Figure 54.—Projected changes in winter mean, minimum, and maximum temperatures across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. S p rin g (M a r - M a y ) 70 T e m p e ra tu re (0 F ) 65 60 55 50 45 40 35 1971 - 2000 2010 - 2039 2040 - 2069 2070 - 2099 GFD L A 1FI Mean G F D L A 1 F I M a xim u m G F D L A 1 F I M in im u m P C M B 1 Mean P C M B 1 M a xim u m P C M B 1 M in im u m Figure 55.—Projected changes in spring mean, minimum, and maximum temperatures across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. 251 APPENDiX 3 S u m m e r (J u n e - Au g ) 95 90 T e m p e ra tu re (0 F ) 85 80 75 70 65 60 55 1971 - 2000 2010 - 2039 2040 - 2069 2070 - 2099 GFD L A 1FI Mean G F D L A 1 F I M a xim u m G F D L A 1 F I M in im u m P C M B 1 Mean P C M B 1 M a xim u m P C M B 1 M in im u m Figure 56.—Projected changes in summer mean, minimum, and maximum temperatures across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. F a ll (S e p - N o v ) 75 70 T e m p e ra tu re (0 F ) 65 60 55 50 45 40 1971 - 2000 2010 - 2039 2040 - 2069 2070 - 2099 GFD L A 1FI Mean G F D L A 1 F I M a xim u m G F D L A 1 F I M in im u m P C M B 1 Mean P C M B 1 M a xim u m P C M B 1 M in im u m Figure 57.—Projected changes in fall mean, minimum, and maximum temperatures across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. 252 APPENDiX 3 S pring (Mar - May) Winter (Dec - F eb) 14 11 P re c ip ita tio n (in c h e s ) P re c ip ita tio n (in c h e s ) 12 10 13 12 9 11 8 1971 - 2000 2010 - 2039 2040 - 2069 GFD L A 1FI 1971 - 2000 2070 - 2099 2010 - 2039 GFD L A 1FI PCM B1 S ummer (J une - Aug) 2040 - 2069 2070 - 2099 PCM B1 F all (S ep - Nov) 15 11 14 P re c ip ita tio n (in c h e s ) P re c ip ita tio n (in c h e s ) 10 13 12 11 10 9 8 9 7 8 1971 - 2000 2010 - 2039 GFD L A 1FI 2040 - 2069 PCM B1 2070 - 2099 1971 - 2000 2010 - 2039 GFD L A 1FI 2040 - 2069 2070 - 2099 PCM B1 Figure 58.—Projected changes in seasonal precipitaion across the assessment area averaged over 30-year periods. The 1971 through 2000 value is based on observed data from weather staions. Note the precipitaion axes are diferent depending on the season. 253 APPENDiX 4: ADDiTioNAL iMPACT MoDEL RESuLTS This appendix provides supplementary information to Chapter 5. The following pages contain additional model results and modifying factors from the Climate Change Tree Atlas, LINKAGES, and LANDIS PRO models. Scientific names for all species are provided in Appendix 1. See Chapter 2 for a description of the models and Chapter 5 for a discussion of model results, uncertainty, and limitations. CLiMATE ChANGE TREE ATLAS MoDEL RESuLTS Tables 29 through 3 show results of the DISTRIB model used in the Tree Atlas averaged over the whole assessment area, and for each section within the assessment area. Section 221E was further divided based on state boundaries into West Virginia 221E and Ohio 221E. Measured area-weighted importance values (IVs) from the U.S. Forest Service, Forest Inventory and Analysis (FIA) as well as modeled current (191 through 1990) and future (2010 through 2039, 2040 through 209, and 2070 through 2099) IVs from DISTRIB were calculated for each time period. One hundred thirty-four tree species were initially modeled. If a species never had an area-weighted IV greater than 3 (FIA, current modeled, or future) across the region, it was deleted from the list because the species either has not had or is not projected to have habitat in the region or there were not enough data. Therefore, only a subset of 93 of the 134 possible species is shown. Three species (blue ash, southern magnolia, and tamarack) were rare within sections and were modeled only at the regional level. Species establishment, growth, and habitat suitability are a function of current (FIA) 254 values. Therefore, it is possible for model results to show species occupying areas where they do not naturally occur (e.g., pine plantations). Conversely, rare species are especially difficult to model at a large regional scale, and may not appear in the FIA data, despite finer inventory data that document their existence. A set of rules was established to determine change classes for the years 2070 through 2099, which was used to create tables in Chapter 5. For most species, the following rules applied, based on the ratio of future IVs to current modeled IVs: Future:Current modeled IV Class <0.5 0.5 through 0.8 >0.8 through <1.2 1.2 through 2.0 >2 large decrease small decrease no change small increase large increase A few exceptions applied to these general rules. When there was a zero in the numerator or denominator, a ratio could not be calculated. Instead, a species was classified as gaining new habitat if its FIA value was 0 and the future IV was greater than 3. A species’ habitat was considered to be extirpated if the future IV was zero and FIA values were greater than 3. Special rules were created for rare species. A species was considered rare if it had a current modeled area-weighted IV that equaled less than 10 percent of the number of pixels in the assessment area (each pixel is a 12.5-mile cell). The change classes are calculated differently for these species because their APPENDiX 4 current infrequency tends to inflate the percentage change that is projected. The cutoffs for each portion of the assessment area were as follows: Section Assessment area 221F OH 221E WV 221E M221A M221B M221C Pixels Cutoff IV for rare species 332 54 85 5 37 53 38 33 5 9 7 4 5 4 When a species was below the cutoff above, it was considered rare, and the following rules applied: Future:Current modeled IV Class <0.2 large decrease 0.2 through <0. small decrease 0. through <4 no change 4 through 8 small increase >8 large increase (not used when current modeled IV ≤3) “Extirpated” was not used in this case because of low confidence. Special rules also applied to species that were known to be present (current FIA IV >0) but not modeled as present (current modeled = 0). In these cases, the FIA IV was used in place of the current modeled IV to calculate ratios. Then, change class rules were applied based on the FIA IV. Tables 37 and 38 describe the modifying factors and adaptability scores used in the Tree Atlas. These factors were developed by using a literature-based scoring system to capture the potential adaptability of species to changes in climate that cannot be adequately captured by DISTRIB (Matthews et al. 2011). This approach was used to assess the capacity for each species to adapt and considered nine biological traits reflecting innate characteristics including competition ability for light and edaphic specificity. Twelve disturbance characteristics addressed the general response of a species to events such as drought, insect pests, and fire. This information distinguishes between species likely to be more tolerant (or sensitive) to environmental changes than the habitat models alone suggest. For each biological and disturbance factor, a species was scored on a scale from -3 through +3. A score of -3 indicated a very negative response of that species to that factor. A score of +3 indicated a very positive response to that factor. To account for confidence in the literature about these factors, each of these scores was then multiplied by 0.5, 0.75, or 1, with 0.5 indicating low confidence and 1 indicating high confidence. Finally the score was further weighted by its relevance to future projected climate change by multiplying it by a relevance factor. A score of 4 indicated highly relevant and 1 indicated not highly relevant to climate change. Means for individual biological scores and disturbance scores were then calculated to arrive at an overall biological and disturbance score for the species. To arrive at an overall adaptability score for the species that could be compared across all modeled tree species, the mean, rescaled (0 through ) values for biological and disturbance characteristics were plotted to form two sides of a right triangle; the hypotenuse was then a combination (disturbance and biological characteristics) metric, ranging from 0 through 8.5 (Fig. 59). Note that modifying factors and adaptability scores are calculated for a species across its entire range. Many species may have higher or lower adaptability in certain areas. For example, a species with a low flooding tolerance may have higher adaptability in areas not subject to flooding. Likewise, local impacts of insects and disease may reduce the adaptability of a species in that area. Only the traits that elicited a combination of a strong positive or negative response, high certainty, and high future relevance for a combined score of 4.5 or greater are listed in the following tables for each species. 255 APPENDiX 4 Mixed hardwoods on the Shawnee State Forest, Ohio. Photo by the Ohio Department of Natural Resources, used with permission. 25 Table 29.—Complete DiSTRiB model resultsa for the 93 tree species in the assessment area Common Name 357 1319 78 1170 13 407 3 17 243 56 23 2163 0 904 87 916 434 63 703 11 8 214 16 71 0 1621 42 19 70 213 58 259 420 294 98 354 1625 88 91 86 3 751 23 2010-2039 PCM GFDL B1 A1Fi 335 1275 59 1102 2 403 7 18 251 44 20 2337 7 900 47 985 485 56 799 8 0 222 5 10 0 1577 42 32 59 173 56 325 485 295 193 446 1571 103 102 72 15 838 4 317 1280 66 1012 5 372 7 12 173 63 6 1864 5 855 57 978 526 40 843 17 6 194 7 37 0 1658 73 11 97 187 31 281 424 357 337 386 1686 97 122 108 9 814 9 Medium High Medium Medium High Medium High Low High Low High High High Low Low High Medium Low High Medium Low Medium Medium Low Low High Medium Low Medium High Low High Medium Medium Medium High High Medium Medium Low High High High 243 948 63 1040 3 365 1 13 51 219 3 1094 293 789 24 1559 559 124 867 333 5 234 15 10 0 1503 270 4 343 175 89 255 459 517 1048 353 1724 206 308 179 47 939 10 296 1275 61 892 5 367 7 10 121 82 3 1577 11 819 54 1018 576 36 899 24 5 211 9 25 0 1703 109 8 113 186 26 278 408 406 464 378 1723 126 149 118 39 846 9 114 647 56 967 2 408 1 14 4 329 4 660 859 660 14 2203 577 218 854 1088 5 293 54 4 239 1082 327 0 571 156 142 210 549 661 1456 217 1440 448 418 282 233 1048 10 DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 294 1272 60 895 5 368 3 12 100 110 2 1481 24 781 49 1037 609 42 910 42 5 209 8 19 0 1622 156 8 128 189 31 276 411 413 590 370 1730 145 167 124 30 847 10 97 547 55 751 2 421 1 14 0 367 5 536 1171 590 14 2118 463 223 854 1473 3 284 70 0 584 979 324 0 810 151 171 196 637 556 1648 196 1141 535 423 350 563 1099 10 0.95 1.00 1.12 0.92 2.50 0.92 1.00 0.67 0.69 1.43 0.30 0.80 0.71 0.95 1.21 0.99 1.09 0.71 1.06 2.13 NA 0.87 1.40 3.70 NA 1.05 1.74 0.34 1.64 1.08 0.55 0.87 0.87 1.21 1.75 0.87 1.07 0.94 1.20 1.50 0.60 0.97 2.25 0.73 0.74 1.07 0.94 1.50 0.91 0.14 0.72 0.20 4.98 0.15 0.47 41.86 0.88 0.51 1.58 1.15 2.21 1.09 41.63 NA 1.05 3.00 1.00 NA 0.95 6.43 0.13 5.81 1.01 1.59 0.79 0.95 1.75 5.43 0.79 1.10 2.00 3.02 2.49 3.13 1.12 2.50 0.88 1.00 1.03 0.81 2.50 0.91 1.00 0.56 0.48 1.86 0.15 0.68 1.57 0.91 1.15 1.03 1.19 0.64 1.13 3.00 NA 0.95 1.80 2.50 NA 1.08 2.60 0.25 1.92 1.08 0.46 0.86 0.84 1.38 2.40 0.85 1.10 1.22 1.46 1.64 2.60 1.01 2.25 0.34 0.51 0.95 0.88 1.00 1.01 0.14 0.78 0.02 7.48 0.20 0.28 122.71 0.73 0.30 2.24 1.19 3.89 1.07 136.00 NA 1.32 10.80 0.40 Migrant 0.69 7.79 0.00 9.68 0.90 2.54 0.65 1.13 2.24 7.54 0.49 0.92 4.35 4.10 3.92 15.53 1.25 2.50 0.88 1.00 1.02 0.81 2.50 0.91 0.43 0.67 0.40 2.50 0.10 0.63 3.43 0.87 1.04 1.05 1.26 0.75 1.14 5.25 NA 0.94 1.60 1.90 NA 1.03 3.71 0.25 2.17 1.09 0.55 0.85 0.85 1.40 3.06 0.83 1.10 1.41 1.64 1.72 2.00 1.01 2.50 0.29 0.43 0.93 0.68 1.00 1.05 0.14 0.78 0.00 8.34 0.25 0.23 167.29 0.66 0.30 2.15 0.96 3.98 1.07 184.13 NA 1.28 14.00 0.00 Migrant 0.62 7.71 0.00 13.73 0.87 3.05 0.60 1.31 1.89 8.54 0.44 0.73 5.19 4.15 4.86 37.53 1.31 2.50 Change Class 2070-2099 PCM GFDL B1 A1Fi No Change No Change No Change No Change No Change No Change Small Decrease No Change Large Decrease Large Increase Large Decrease Small Decrease No Change No Change No Change No Change Small Increase Small Decrease No Change Small Increase No Change No Change No Change No Change NA No Change Large Increase Small Decrease Large Increase No Change Small Decrease No Change No Change Small Increase Large Increase No Change No Change Small Increase Small Increase Small Increase No Change No Change No Change Large Decrease Large Decrease No Change Small Decrease No Change No Change Large Decrease No Change Exirpated Large Increase Small Decrease Large Decrease Large Increase Small Decrease Large Decrease Large Increase No Change Large Increase No Change Large Increase Small Decrease Small Increase Large Increase Exirpated New Habitat Small Decrease Large Increase Exirpated Large Increase No Change Large Increase Small Decrease Small Increase Small Increase Large Increase Large Decrease Small Decrease Large Increase Large Increase Large Increase Large Increase Small Increase No Change 257 (coninued on next page) APPENDiX 4 American basswood American beech American chestnut American elm American holly American hornbeam Balsam ir Bear oak (scrub oak) Bigtooth aspen Biternut hickory Black ash Black cherry Black hickory Black locust Black maple Black oak Black walnut Black willow Blackgum Blackjack oak Blue ash Boxelder Bur oak Buternut Cedar elm Chestnut oak Chinkapin oak Chokecherry Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redbud Eastern redcedar Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Mountain maple FiA iV DiSTRiB Current Model iV Reliability Modeled iV 2040-2069 PCM GFDL B1 A1Fi Common Name Northern catalpa Northern pin oak Northern red oak Northern white-cedar Ohio buckeye Osage-orange Pawpaw Pignut hickory Pin cherry Pin oak Pitch pine Post oak Quaking aspen Red maple Red mulberry Red pine Red spruce River birch Rock elm Sassafras Scarlet oak Serviceberry Shagbark hickory Shingle oak Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern magnolia Southern red oak Striped maple Sugar maple Sugarberry Swamp white oak Sweet birch Sweetgum Sycamore Table Mountain pine Tamarack (naive) Tulip tree Virginia pine Water locust FiA iV 12 6 1397 0 51 63 172 759 52 128 134 47 54 3395 13 74 67 20 5 1111 514 287 144 31 40 1 183 610 312 2 11 261 2817 1 59 440 15 277 44 3 1685 489 0 DiSTRiB Current Model iV Reliability 2010-2039 PCM GFDL B1 A1Fi 4 2 1400 4 32 67 132 864 18 87 138 57 82 3501 13 51 39 4 1 1143 582 259 207 8 53 0 135 596 362 4 5 226 2601 1 53 474 8 277 41 2 1652 526 0 3 4 1313 1 37 67 133 848 15 79 131 120 36 3240 13 42 40 11 4 1179 612 269 204 22 79 0 96 590 400 4 14 206 2700 0 42 453 26 305 39 1 1783 581 0 Low Medium High High Low Medium Low High Medium Medium High High High High Low Medium High Low Low High High Medium Medium Medium High Low Medium Medium High Medium High High High Medium Low High High Medium Medium High High High Low 3 3 1322 1 76 113 132 940 3 131 129 1513 5 2390 129 20 20 14 11 1154 705 239 408 125 467 7 222 617 336 1 136 142 2226 57 51 353 85 346 46 1 1564 579 0 Modeled iV 2040-2069 PCM GFDL B1 A1Fi 3 3 1260 1 41 86 146 874 10 80 131 176 22 3179 15 30 33 12 3 1191 657 251 249 19 115 0 87 558 505 8 43 179 2681 5 41 453 51 310 39 1 1895 716 0 7 1 1206 5 21 159 69 810 2 197 121 3340 4 1664 295 10 15 16 44 880 522 227 558 123 1312 132 407 529 275 2 374 103 1068 328 26 248 222 352 53 6 939 500 6 DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 2 3 1244 1 49 92 144 887 7 79 124 257 14 3114 23 24 34 10 4 1229 646 243 277 35 152 0 98 560 480 4 54 170 2700 9 43 451 78 317 36 1 1939 700 0 20 1 1099 6 5 254 51 766 2 158 129 4767 3 1299 366 10 14 32 20 820 345 230 559 106 1972 282 425 402 223 2 609 95 587 648 6 208 357 368 63 6 687 491 56 0.75 2.00 0.94 0.25 1.16 1.00 1.01 0.98 0.83 0.91 0.95 2.11 0.44 0.93 1.00 0.82 1.03 2.75 4.00 1.03 1.05 1.04 0.99 2.75 1.49 NA 0.71 0.99 1.11 1.00 2.80 0.91 1.04 0.00 0.79 0.96 3.25 1.10 0.95 0.50 1.08 1.11 NA 0.75 1.50 0.94 0.25 2.38 1.69 1.00 1.09 0.17 1.51 0.94 26.54 0.06 0.68 9.92 0.39 0.51 3.50 11.00 1.01 1.21 0.92 1.97 15.63 8.81 NA 1.64 1.04 0.93 0.25 27.20 0.63 0.86 57.00 0.96 0.75 10.63 1.25 1.12 0.50 0.95 1.10 NA 0.75 1.50 0.90 0.25 1.28 1.28 1.11 1.01 0.56 0.92 0.95 3.09 0.27 0.91 1.15 0.59 0.85 3.00 3.00 1.04 1.13 0.97 1.20 2.38 2.17 NA 0.64 0.94 1.40 2.00 8.60 0.79 1.03 5.00 0.77 0.96 6.38 1.12 0.95 0.50 1.15 1.36 NA 1.75 0.50 5.00 0.50 1.50 0.50 0.86 0.89 0.79 1.25 0.25 1.50 0.66 1.53 0.16 2.37 1.37 3.79 0.52 1.09 0.39 0.94 1.03 0.89 0.11 0.39 0.11 2.26 0.91 1.82 0.88 0.90 0.94 58.60 4.51 83.63 0.05 0.17 0.04 0.48 0.89 0.37 22.69 1.77 28.15 0.20 0.47 0.20 0.39 0.87 0.36 4.00 2.50 8.00 44.00 4.00 20.00 0.77 1.08 0.72 0.90 1.11 0.59 0.88 0.94 0.89 2.70 1.34 2.70 15.38 4.38 13.25 24.76 2.87 37.21 NA NA NA 3.02 0.73 3.15 0.89 0.94 0.67 0.76 1.33 0.62 0.50 1.00 0.50 74.80 10.80 121.80 0.46 0.75 0.42 0.41 1.04 0.23 328.00 9.00 648.00 0.49 0.81 0.11 0.52 0.95 0.44 27.75 9.75 44.63 1.27 1.14 1.33 1.29 0.88 1.54 3.00 0.50 3.00 0.57 1.17 0.42 0.95 1.33 0.93 Migrant NA Migrant Change Class 2070-2099 PCM GFDL B1 A1Fi No Change No Change No Change Small Decrease No Change Small Increase No Change No Change Small Decrease No Change No Change Large Increase Large Decrease No Change No Change Large Decrease No Change No Change No Change No Change No Change No Change Small Increase Small Increase Large Increase NA Small Decrease No Change Small Increase No Change Large Increase Small Decrease No Change Large Increase No Change No Change Large Increase No Change No Change No Change No Change Small Increase NA Small Increase No Change Small Decrease No Change Large Decrease Large Increase Large Decrease No Change Large Decrease Small Increase No Change Large Increase Large Decrease Large Decrease Large Increase Large Decrease Large Decrease Small Increase Large Increase Small Decrease Small Decrease No Change Large Increase Large Increase Large Increase Large Increase Large Increase Small Decrease Small Decrease No Change Large Increase Large Decrease Large Decrease Large Increase Large Decrease Large Decrease Large Increase Small Increase Small Increase No Change Large Decrease No Change New Habitat (coninued on next page) APPENDiX 4 258 Table 29 (coninued). Table 29 (coninued). Common Name FiA iV Water oak White ash White oak Willow oak Winged elm Yellow birch Yellow buckeye 0 1483 1572 2 3 151 202 DiSTRiB Current Model iV Reliability 2010-2039 PCM GFDL B1 A1Fi 0 1657 1643 0 2 135 161 0 1477 1720 0 11 113 176 High High High Medium High High Medium 0 1250 2487 0 388 53 175 Modeled iV 2040-2069 PCM GFDL B1 A1Fi 0 1394 1791 0 43 106 175 89 827 2519 1 1229 33 177 DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 0 1422 1853 0 68 103 176 323 747 2023 20 2272 29 174 NA 0.89 1.05 NA 5.50 0.84 1.09 Change Class 2070-2099 PCM GFDL B1 A1Fi NA NA Migrant NA Migrant NA 0.75 0.84 0.50 0.86 0.45 No Change 1.51 1.09 1.53 1.13 1.23 No Change NA NA NA NA NA NA 194.00 21.50 614.50 34.00 1136.00 Large Increase 0.39 0.79 0.24 0.76 0.22 Small Decrease 1.09 1.09 1.10 1.09 1.08 No Change New Habitat Large Decrease Small Increase Large Increase Large Increase Large Decrease No Change Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 259 Common Name FiA iV American basswood American beech American elm American hornbeam Bigtooth aspen Biternut hickory Black ash Black cherry Black hickory Black locust Black maple Black oak Black walnut Black willow Blackgum Blackjack oak Boxelder Bur oak Cedar elm Chestnut oak Chinkapin oak Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Mockernut hickory Northern pin oak Northern red oak Northern white-cedar Ohio buckeye Osage-orange Pignut hickory Pin cherry Pin oak Post oak 73 228 495 79 34 8 7 705 0 74 30 40 55 30 35 0 31 12 0 0 0 0 18 26 9 91 0 0 48 115 16 4 28 84 4 191 0 4 45 86 15 106 0 DiSTRiB Current Model iV Reliability 73 178 454 73 57 22 20 691 0 101 17 79 80 36 34 0 51 5 0 12 1 0 5 34 18 123 6 10 44 113 29 21 19 101 2 192 3 7 26 107 6 74 0 Medium High Medium Medium High Low High High High Low Low High Medium Low High Medium Medium Medium Low High Medium Medium High Low High Medium Medium Medium High High Medium Medium Low High Medium High High Low Medium High Medium Medium High 2010-2039 PCM GFDL B1 A1Fi 56 182 468 68 53 20 5 655 0 115 18 69 108 28 50 0 44 6 0 29 4 0 11 20 9 82 29 30 50 146 21 34 47 93 4 168 0 12 39 105 5 68 0 55 151 434 62 18 59 3 289 7 114 10 132 137 49 33 8 52 13 0 34 56 27 7 53 3 98 173 77 40 143 42 125 76 140 3 191 0 39 60 152 2 89 112 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 56 180 426 68 55 37 2 606 0 123 18 91 152 26 59 2 49 8 0 46 13 3 11 16 8 75 79 50 51 155 38 51 62 106 3 169 0 17 52 115 5 65 9 57 183 422 68 44 47 1 546 0 121 18 96 162 32 59 3 49 7 0 42 30 1 11 20 5 73 111 57 41 158 40 57 66 104 3 166 0 24 52 114 4 63 15 11 62 403 69 0 64 4 111 56 72 5 237 143 61 32 120 64 42 2 6 64 45 5 79 1 64 218 159 0 116 135 152 97 170 1 178 5 16 66 159 1 107 401 5 51 224 67 0 58 5 41 105 99 5 223 103 57 38 149 59 43 61 3 54 110 5 76 1 83 220 114 0 115 146 133 104 153 1 116 6 4 79 130 1 84 655 0.77 1.02 1.03 0.93 0.93 0.91 0.25 0.95 NA 1.14 1.06 0.87 1.35 0.78 1.47 NA 0.86 1.20 NA 2.42 4.00 NA 2.20 0.59 0.50 0.67 4.83 3.00 1.14 1.29 0.72 1.62 2.47 0.92 2.00 0.88 0.00 1.71 1.50 0.98 0.83 0.92 NA 0.75 0.85 0.96 0.85 0.32 2.68 0.15 0.42 New 1.13 0.59 1.67 1.71 1.36 0.97 New 1.02 2.60 NA 2.83 56.00 New 1.40 1.56 0.17 0.80 28.83 7.70 0.91 1.27 1.45 5.95 4.00 1.39 1.50 1.00 0.00 5.57 2.31 1.42 0.33 1.20 New 0.77 0.15 1.01 0.35 0.94 0.89 0.93 0.95 0.97 0.00 1.68 2.91 0.10 0.20 0.88 0.16 NA New 1.22 0.71 1.06 0.29 1.15 3.00 1.90 1.79 0.72 1.69 1.74 0.94 New New 0.96 1.26 1.60 8.40 NA New 3.83 0.50 13.00 64.00 New New 2.20 1.00 0.47 2.32 0.44 0.06 0.61 0.52 13.17 36.33 5.00 15.90 1.16 0.00 1.37 1.03 1.31 4.66 2.43 7.24 3.26 5.11 1.05 1.68 1.50 0.50 0.88 0.93 0.00 1.67 2.43 2.29 2.00 2.54 1.08 1.49 0.83 0.17 0.88 1.45 New New 0.78 0.07 1.03 0.29 0.93 0.49 0.93 0.92 0.77 0.00 2.14 2.64 0.05 0.25 0.79 0.06 NA New 1.20 0.98 1.06 0.29 1.22 2.82 2.03 1.29 0.89 1.58 1.74 1.12 New New 0.96 1.16 1.40 8.60 NA New 3.50 0.25 30.00 54.00 New New 2.20 1.00 0.59 2.24 0.28 0.06 0.59 0.68 18.50 36.67 5.70 11.40 0.93 0.00 1.40 1.02 1.38 5.03 2.71 6.33 3.47 5.47 1.03 1.52 1.50 0.50 0.87 0.60 0.00 2.00 3.43 0.57 2.00 3.04 1.07 1.22 0.67 0.17 0.85 1.14 New New Change Class 2070-2099 PCM GFDL B1 A1Fi Decrease No Change No Change No Change Decrease Large Increase Large Decrease Decrease NA Increase No Change Increase Increase No Change Increase New Habitat No Change Increase NA New Habitat New Habitat New Habitat Large Increase Decrease Large Decrease Decrease New Habitat New Habitat No Change Increase Increase Large Increase Large Increase No Change No Change No Change NA Large Increase Increase No Change Decrease No Change New Habitat Large Decrease Large Decrease Decrease No Change Large Decrease Large Increase Large Decrease Large Decrease New Habitat No Change Large Decrease Large Increase Increase Increase No Change New Habitat Increase Large Increase New Habitat Large Decrease New Habitat New Habitat No Change Large Increase Large Decrease Decrease New Habitat New Habitat Large Decrease No Change Large Increase Large Increase Large Increase Increase Decrease Decrease No Change Decrease Large Increase Increase Large Decrease No Change New Habitat (coninued on next page) APPENDiX 4 20 Table 30.—Complete DiSTRiB model resultsa for the 68 tree species in Secion 221F Table 30 (coninued). Common Name FiA iV Quaking aspen Red maple Red mulberry Red pine Rock elm Sassafras Scarlet oak Serviceberry Shagbark hickory Shingle oak Shortleaf pine Shumard oak Silver maple Slippery elm Southern red oak Sugar maple Sugarberry Swamp white oak Sweet birch Sweetgum Sycamore Tulip tree White oak Winged elm Yellow birch 33 553 0 29 0 32 7 3 25 7 0 0 90 101 0 538 0 52 1 0 22 66 53 0 7 DiSTRiB Current Model iV Reliability 49 537 1 28 0 72 6 2 44 0 0 0 56 121 0 490 0 48 3 0 41 75 118 0 9 High High Low Medium Low High High Medium Medium Medium High Low Medium Medium High High Medium Low High High Medium High High High High 2010-2039 PCM GFDL B1 A1Fi 23 418 1 26 0 91 8 2 36 5 0 0 45 139 0 509 0 41 4 0 60 85 103 0 6 0 288 49 11 0 102 21 1 109 69 0 0 95 149 0 480 0 46 2 0 63 80 230 0 0 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 13 390 4 17 0 112 29 2 61 5 0 0 43 146 0 544 0 39 9 2 63 118 125 0 3 7 373 7 16 0 119 28 2 69 7 0 0 49 145 0 562 0 41 11 3 64 124 134 0 1 3 237 61 9 9 80 38 1 131 67 8 0 141 141 1 191 14 25 0 1 56 59 260 30 0 3 201 78 9 11 97 30 1 79 49 40 27 104 77 14 109 121 6 0 6 61 52 167 217 0 0.47 0.78 1.00 0.93 NA 1.26 1.33 1.00 0.82 New NA NA 0.80 1.15 NA 1.04 NA 0.85 1.33 NA 1.46 1.13 0.87 NA 0.67 0.00 0.54 49.00 0.39 NA 1.42 3.50 0.50 2.48 New NA NA 1.70 1.23 NA 0.98 NA 0.96 0.67 NA 1.54 1.07 1.95 NA 0.00 0.27 0.06 0.73 0.44 4.00 61.00 0.61 0.32 NA New 1.56 1.11 4.83 6.33 1.00 0.50 1.39 2.98 New New NA New NA NA 0.77 2.52 1.21 1.17 NA New 1.11 0.39 NA New 0.81 0.52 3.00 0.00 New New 1.54 1.37 1.57 0.79 1.06 2.20 NA New 0.33 0.00 0.14 0.06 0.70 0.37 7.00 78.00 0.57 0.32 New New 1.65 1.35 4.67 5.00 1.00 0.50 1.57 1.80 New New New New New New 0.88 1.86 1.20 0.64 New New 1.15 0.22 New New 0.85 0.13 3.67 0.00 New New 1.56 1.49 1.65 0.69 1.14 1.42 New New 0.11 0.00 Change Class 2070-2099 PCM GFDL B1 A1Fi Large Decrease Decrease New Habitat Decrease NA Increase Large Increase No Change Increase No Change NA NA No Change Increase NA No Change NA No Change No Change New Habitat Increase Increase No Change NA Large Decrease Large Decrease Large Decrease New Habitat Large Decrease New Habitat Increase Large Increase Decrease Increase Increase New Habitat New Habitat Increase Decrease New Habitat Large Decrease New Habitat Large Decrease NA New Habitat Increase Decrease Increase New Habitat Large Decrease Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 21 Common Name FiA iV American basswood American beech American elm American hornbeam Bigtooth aspen Biternut hickory Black cherry Black hickory Black locust Black maple Black oak Black walnut Black willow Blackgum Blackjack oak Boxelder Bur oak Cedar elm Chestnut oak Chinkapin oak Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Northern catalpa Northern red oak Ohio buckeye Osage-orange Pawpaw Pignut hickory Pin oak Pitch pine Post oak 43 194 395 133 149 15 582 0 259 41 248 165 15 133 2 47 2 0 220 13 32 12 27 9 106 47 108 98 595 24 30 40 0 198 9 233 29 12 51 204 16 24 16 DiSTRiB Current Model iV Reliability 55 202 379 108 116 12 650 0 254 23 261 171 13 144 1 62 0 0 219 16 31 2 9 13 120 73 94 98 476 25 39 37 1 206 1 271 15 26 48 218 8 16 17 Medium High Medium Medium High Low High High Low Low High Medium Low High Medium Medium Medium Low High Medium Medium High Low High Medium Medium Medium High High Medium Medium Low High High Low High Low Medium Low High Medium High High 2010-2039 PCM GFDL B1 A1Fi 45 212 305 114 72 17 459 0 203 27 288 169 5 174 4 45 0 0 248 32 46 7 9 3 103 141 119 75 552 26 39 41 0 220 2 245 13 18 41 239 7 20 40 38 149 316 112 11 76 223 115 179 10 508 166 46 179 130 65 0 0 242 90 134 5 28 3 110 358 164 49 517 73 101 67 4 245 2 294 27 30 52 265 29 15 658 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 49 217 254 111 38 20 348 2 180 25 301 172 4 193 7 52 0 0 278 50 52 7 6 3 100 182 131 58 546 31 47 37 2 231 2 249 11 21 48 251 9 23 60 49 228 258 113 35 22 335 6 165 25 303 175 4 188 11 49 0 0 256 62 53 8 5 4 103 219 128 55 548 37 55 39 1 230 1 253 13 25 48 259 9 21 86 6 78 249 122 0 87 155 224 158 7 640 151 73 151 339 104 8 107 174 87 142 4 46 2 156 390 189 1 355 163 118 94 37 256 1 229 2 48 30 234 68 15 1052 3 62 177 127 0 95 137 308 128 7 566 93 63 153 441 98 21 194 167 76 188 4 67 2 169 436 136 1 265 180 106 126 109 258 1 220 1 93 18 231 51 14 1388 0.82 1.05 0.81 1.06 0.62 1.42 0.71 NA 0.80 1.17 1.10 0.99 0.39 1.21 4.00 0.73 NA NA 1.13 2.00 1.48 3.50 1.00 0.23 0.86 1.93 1.27 0.77 1.16 1.04 1.00 1.11 0.00 1.07 2.00 0.90 0.87 0.69 0.85 1.10 0.88 1.25 2.35 0.69 0.74 0.83 1.04 0.10 6.33 0.34 New 0.71 0.44 1.95 0.97 3.54 1.24 130.00 1.05 NA NA 1.11 5.63 4.32 2.50 3.11 0.23 0.92 4.90 1.75 0.50 1.09 2.92 2.59 1.81 4.00 1.19 2.00 1.09 1.80 1.15 1.08 1.22 3.63 0.94 38.71 0.89 0.11 1.07 0.39 0.67 0.66 1.03 1.13 0.33 0.00 1.67 7.25 0.54 0.24 New New 0.71 0.62 1.09 0.30 1.15 2.45 1.01 0.88 0.31 5.62 1.34 1.05 7.00 339.00 0.84 1.68 NA New NA New 1.27 0.80 3.13 5.44 1.68 4.58 3.50 2.00 0.67 5.11 0.23 0.15 0.83 1.30 2.49 5.34 1.39 2.01 0.59 0.01 1.15 0.75 1.24 6.52 1.21 3.03 1.00 2.54 2.00 37.00 1.12 1.24 2.00 1.00 0.92 0.85 0.73 0.13 0.81 1.85 1.00 0.63 1.15 1.07 1.13 8.50 1.44 0.94 3.53 61.88 0.89 0.06 1.13 0.31 0.68 0.47 1.05 1.18 0.30 0.00 1.83 7.92 0.52 0.21 New New 0.65 0.50 1.09 0.30 1.16 2.17 1.02 0.54 0.31 4.85 1.31 1.06 11.00 441.00 0.79 1.58 New New New New 1.17 0.76 3.88 4.75 1.71 6.07 4.00 2.00 0.56 7.44 0.31 0.15 0.86 1.41 3.00 5.97 1.36 1.45 0.56 0.01 1.15 0.56 1.48 7.20 1.41 2.72 1.05 3.41 1.00 109.00 1.12 1.25 1.00 1.00 0.93 0.81 0.87 0.07 0.96 3.58 1.00 0.38 1.19 1.06 1.13 6.38 1.31 0.88 5.06 81.65 Change Class 2070-2099 PCM GFDL B1 A1Fi No Change No Change Decrease No Change Large Decrease Increase Decrease New Habitat Decrease No Change Increase No Change Large Decrease Increase Increase Decrease Large Decrease NA Increase Large Increase Increase Large Increase Decrease Large Decrease No Change Large Increase Increase Decrease Increase Increase Increase No Change NA No Change No Change No Change No Change No Change No Change Increase No Change Increase Large Increase Large Decrease Large Decrease Decrease Increase Exirpated Large Increase Large Decrease New Habitat Decrease Large Decrease Large Increase Decrease Large Increase No Change Increase Increase Large Increase New Habitat Decrease Large Increase Large Increase Increase Large Increase Large Decrease Increase Large Increase Increase Exirpated Decrease Large Increase Large Increase Large Increase New Habitat Increase No Change Decrease Large Decrease Large Increase Large Decrease No Change Large Increase No Change Large Increase (coninued on next page) APPENDiX 4 22 Table 31.—Complete DiSTRiB model resultsa for the 72 tree species in Secion 221E (Ohio) Table 31 (coninued). Common Name FiA iV Quaking aspen Red maple Red mulberry Red pine River birch Sassafras Scarlet oak Serviceberry Shagbark hickory Shingle oak Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern red oak Sugar maple Sugarberry Sweet birch Sweetgum Sycamore Tulip tree Virginia pine Water locust Water oak White ash White oak Winged elm Yellow buckeye 12 816 0 18 10 323 88 33 49 19 15 0 57 182 69 0 665 0 8 10 111 473 117 0 0 409 411 0 57 DiSTRiB Current Model iV Reliability 10 769 2 6 3 323 108 25 76 8 15 0 52 195 49 1 634 1 26 1 109 429 126 0 0 444 429 1 47 High High Low Medium Low High High Medium Medium Medium High Low Medium Medium High High High Medium High High Medium High High Low High High High High Medium 2010-2039 PCM GFDL B1 A1Fi 3 734 2 3 5 341 127 30 75 17 26 0 29 175 81 3 652 0 15 6 115 484 173 0 0 391 482 1 49 0 397 41 1 10 328 165 29 143 48 136 3 81 198 43 31 524 24 6 11 119 342 130 0 0 323 784 159 46 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 1 718 2 2 7 355 135 28 86 13 35 0 23 167 109 8 651 1 23 9 115 499 222 0 0 364 518 7 49 0 690 2 0 7 365 132 26 94 22 44 0 27 168 89 10 659 2 22 14 113 501 210 0 0 373 532 13 48 0 265 87 1 13 239 106 29 146 46 299 41 185 143 34 83 163 123 2 25 103 195 97 1 23 202 655 342 50 0 227 93 1 22 218 74 27 128 45 445 89 187 95 36 137 76 226 2 47 95 159 106 16 87 174 472 682 48 0.30 0.95 1.00 0.50 1.67 1.06 1.18 1.20 0.99 2.13 1.73 NA 0.56 0.90 1.65 3.00 1.03 0.00 0.58 6.00 1.06 1.13 1.37 NA NA 0.88 1.12 1.00 1.04 0.00 0.52 20.50 0.17 3.33 1.02 1.53 1.16 1.88 6.00 9.07 New 1.56 1.02 0.88 31.00 0.83 24.00 0.23 11.00 1.09 0.80 1.03 NA NA 0.73 1.83 159.00 0.98 0.10 0.93 1.00 0.33 2.33 1.10 1.25 1.12 1.13 1.63 2.33 NA 0.44 0.86 2.22 8.00 1.03 1.00 0.89 9.00 1.06 1.16 1.76 NA NA 0.82 1.21 7.00 1.04 0.00 0.35 43.50 0.17 4.33 0.74 0.98 1.16 1.92 5.75 19.93 New 3.56 0.73 0.69 83.00 0.26 123.00 0.08 25.00 0.95 0.46 0.77 New New 0.46 1.53 342.00 1.06 0.00 0.90 1.00 0.00 2.33 1.13 1.22 1.04 1.24 2.75 2.93 New 0.52 0.86 1.82 10.00 1.04 2.00 0.85 14.00 1.04 1.17 1.67 New New 0.84 1.24 13.00 1.02 0.00 0.30 46.50 0.17 7.33 0.68 0.69 1.08 1.68 5.63 29.67 New 3.60 0.49 0.74 137.00 0.12 226.00 0.08 47.00 0.87 0.37 0.84 New New 0.39 1.10 682.00 1.02 Change Class 2070-2099 PCM GFDL B1 A1Fi Exirpated No Change NA Large Decrease No Change No Change Increase No Change Increase No Change Large Increase NA Decrease No Change Increase New Habitat No Change New Habitat Decrease Increase No Change Increase Increase NA NA Decrease Increase New Habitat No Change Exirpated Large Decrease New Habitat Large Decrease Increase Decrease Decrease No Change Increase Increase Large Increase NA Decrease No Change Increase New Habitat No Change New Habitat Decrease Increase No Change Increase Increase NA NA Decrease Increase New Habitat No Change Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 23 Common Name FiA iV American basswood American beech American chestnut American elm American hornbeam Bigtooth aspen Biternut hickory Black cherry Black hickory Black locust Black maple Black oak Black walnut Black willow Blackgum Blackjack oak Boxelder Buternut Cedar elm Chestnut oak Chinkapin oak Common persimmon Cucumbertree Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Northern red oak Ohio buckeye Osage-orange Pawpaw Pignut hickory Pin cherry Pin oak Pitch pine Post oak 58 215 7 143 110 32 11 302 0 170 7 208 95 4 117 2 51 12 0 206 5 28 40 12 85 7 93 14 379 23 23 4 2 192 218 3 4 60 191 10 0 21 20 DiSTRiB Current Model iV Reliability 53 211 1 141 97 38 5 333 4 142 4 224 103 0 137 3 57 1 0 235 11 23 30 15 84 45 94 28 407 18 14 4 1 211 211 2 8 49 208 2 2 20 27 Medium High Medium Medium Medium High Low High High Low Low High Medium Low High Medium Medium Low Low High Medium Medium High High Medium Medium Medium High High Medium Medium Low High High High Low Medium Low High Medium Medium High High 2010-2039 PCM GFDL B1 A1Fi 59 233 2 116 100 22 10 220 2 129 4 209 117 0 141 4 54 6 0 238 18 32 39 13 91 85 105 24 402 24 19 3 3 203 198 1 4 53 208 1 1 20 43 24 151 2 129 89 2 45 121 89 111 1 337 99 12 148 113 60 0 0 170 62 76 34 8 92 221 119 23 391 47 36 14 21 218 196 0 13 48 203 0 3 11 390 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 50 232 2 100 99 13 8 161 3 118 4 209 110 0 152 4 58 5 0 258 24 30 39 15 92 101 114 29 401 28 21 3 17 206 194 1 5 53 207 0 1 19 42 43 237 2 101 97 6 21 154 11 111 2 213 124 0 158 16 60 3 0 243 39 33 39 15 92 130 112 40 403 38 22 5 11 211 194 1 6 53 208 0 3 16 73 4 83 0 115 84 0 62 101 268 96 1 495 102 35 151 300 67 0 82 114 64 124 29 3 115 273 114 3 274 75 60 45 80 229 195 0 28 23 131 0 10 8 858 3 76 0 120 87 0 74 100 302 82 1 441 83 42 149 356 63 0 150 115 64 137 29 4 129 298 100 1 175 88 56 54 169 245 207 0 54 18 119 0 15 8 1089 1.11 1.10 2.00 0.82 1.03 0.58 2.00 0.66 0.50 0.91 1.00 0.93 1.14 NA 1.03 1.33 0.95 6.00 NA 1.01 1.64 1.39 1.30 0.87 1.08 1.89 1.12 0.86 0.99 1.33 1.36 0.75 3.00 0.96 0.94 0.50 0.50 1.08 1.00 0.50 0.50 1.00 1.59 0.45 0.72 2.00 0.92 0.92 0.05 9.00 0.36 22.25 0.78 0.25 1.50 0.96 New 1.08 37.67 1.05 0.00 NA 0.72 5.64 3.30 1.13 0.53 1.10 4.91 1.27 0.82 0.96 2.61 2.57 3.50 21.00 1.03 0.93 0.00 1.63 0.98 0.98 0.00 1.50 0.55 14.44 0.94 1.10 2.00 0.71 1.02 0.34 1.60 0.48 0.75 0.83 1.00 0.93 1.07 New 1.11 1.33 1.02 5.00 NA 1.10 2.18 1.30 1.30 1.00 1.10 2.24 1.21 1.04 0.99 1.56 1.50 0.75 17.00 0.98 0.92 0.50 0.63 1.08 1.00 0.00 0.50 0.95 1.56 0.08 0.39 0.00 0.82 0.87 0.00 12.40 0.30 67.00 0.68 0.25 2.21 0.99 New 1.10 100.00 1.18 0.00 New 0.49 5.82 5.39 0.97 0.20 1.37 6.07 1.21 0.11 0.67 4.17 4.29 11.25 80.00 1.09 0.92 0.00 3.50 0.47 0.63 0.00 5.00 0.40 31.78 0.81 1.12 2.00 0.72 1.00 0.16 4.20 0.46 2.75 0.78 0.50 0.95 1.20 New 1.15 5.33 1.05 3.00 New 1.03 3.55 1.44 1.30 1.00 1.10 2.89 1.19 1.43 0.99 2.11 1.57 1.25 11.00 1.00 0.92 0.50 0.75 1.08 1.00 0.00 1.50 0.80 2.70 0.06 0.36 0.00 0.85 0.90 0.00 14.80 0.30 75.50 0.58 0.25 1.97 0.81 New 1.09 118.67 1.11 0.00 New 0.49 5.82 5.96 0.97 0.27 1.54 6.62 1.06 0.04 0.43 4.89 4.00 13.50 169.00 1.16 0.98 0.00 6.75 0.37 0.57 0.00 7.50 0.40 40.33 Change Class 2070-2099 PCM GFDL B1 A1Fi Decrease Decrease No Change Decrease Decrease Large Decrease Increase Large Decrease New Habitat Decrease Decrease Decrease Increase NA Decrease Increase Decrease No Change NA Decrease No Change Increase Increase Decrease Decrease Large Increase Decrease Increase Decrease Large Increase Increase No change Increase Decrease Decrease Decrease No Change Decrease Decrease Exirpated New Habitat Decrease Large Increase Large Decrease Large Decrease Exirpated No Change No Change Exirpated Large Increase Large Decrease New Habitat Decrease Decrease Increase Decrease Increase No Change Increase No Change Exirpated New Habitat Decrease Increase Large Increase No Change Large Decrease Increase Large Increase No Change Exirpated Large Decrease Large Increase Large Increase Large Increase Increase Increase No Change Exirpated Increase Large Decrease Decrease Exirpated New Habitat Large Decrease Large Increase (coninued on next page) APPENDiX 4 24 Table 32.—Complete DiSTRiB model results for the 71 tree speciesa in Secion 221E (West Virginia) Table 32 (coninued). Common Name FiA iV Red maple Red mulberry Rock elm Sassafras Scarlet oak Serviceberry Shagbark hickory Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern red oak Sugar maple Sugarberry Sweet birch Sweetgum Sycamore Table Mountain pine Tulip tree Virginia pine Water locust Water oak White ash White oak Willow oak Winged elm Yellow buckeye 588 5 1 317 105 32 30 6 0 9 191 75 4 563 0 36 2 68 7 454 174 0 0 341 423 0 2 77 DiSTRiB Current Model iV Reliability 599 4 1 305 121 31 46 13 0 6 158 90 2 487 0 38 1 66 2 428 168 0 0 331 411 0 0 56 High Low Low High High Medium Medium High Low Medium Medium High High High Medium High High Medium Medium High High Low High High High Medium High Medium 2010-2039 PCM GFDL B1 A1Fi 569 4 3 315 127 32 48 21 0 3 148 97 4 534 0 41 7 65 3 457 159 0 0 292 422 0 5 64 360 15 3 267 153 26 65 184 3 8 124 58 57 389 19 15 29 66 4 367 149 0 0 210 596 0 127 61 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 564 4 2 298 132 31 51 40 0 3 128 121 19 516 1 42 19 65 2 462 201 0 0 270 432 0 21 61 546 4 3 303 135 33 57 64 0 2 127 116 28 516 3 44 31 66 2 463 202 0 0 280 452 0 33 63 199 58 10 175 60 28 98 459 50 12 82 31 144 115 84 5 66 67 4 126 74 2 31 135 566 0 416 57 173 63 3 157 28 32 113 625 66 43 67 13 204 46 120 5 91 66 4 100 71 16 102 129 427 9 598 57 0.95 1.00 3.00 1.03 1.05 1.03 1.04 1.62 NA 0.50 0.94 1.08 2.00 1.10 NA 1.08 7.00 0.99 1.50 1.07 0.95 NA NA 0.88 1.03 NA New 1.14 0.60 3.75 3.00 0.88 1.26 0.84 1.41 14.15 New 1.33 0.79 0.64 28.50 0.80 New 0.40 29.00 1.00 2.00 0.86 0.89 NA NA 0.63 1.45 NA New 1.09 0.94 1.00 2.00 0.98 1.09 1.00 1.11 3.08 NA 0.50 0.81 1.34 9.50 1.06 New 1.11 19.00 0.99 1.00 1.08 1.20 NA NA 0.82 1.05 NA New 1.09 0.33 14.50 10.00 0.57 0.50 0.90 2.13 35.31 New 2.00 0.52 0.34 72.00 0.24 New 0.13 66.00 1.02 2.00 0.29 0.44 New New 0.41 1.38 NA New 1.02 0.91 1.00 3.00 0.99 1.12 1.07 1.24 4.92 New 0.33 0.80 1.29 14.00 1.06 New 1.16 31.00 1.00 1.00 1.08 1.20 New New 0.85 1.10 New New 1.13 0.29 15.75 3.00 0.52 0.23 1.03 2.46 48.08 New 7.17 0.42 0.14 102.00 0.09 New 0.13 91.00 1.00 2.00 0.23 0.42 New New 0.39 1.04 New New 1.02 Change Class 2070-2099 PCM GFDL B1 A1Fi Decrease No Change Large Increase Decrease Decrease Decrease Increase Large Increase NA Large Decrease Decrease Increase Increase Decrease New Habitat Decrease Increase Decrease No Change Decrease Increase NA NA Decrease Decrease NA Increase Decrease Large Decrease Large Increase Large Increase Decrease Large Decrease No Change Large Increase Large Increase New Habitat Large Increase Large Decrease Large Decrease Increase Large Decrease New Habitat Large Decrease Increase No Change No Change Large Decrease Large Decrease New Habitat New Habitat Large Decrease No Change New Habitat Increase No Change Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 25 Common Name FiA iV American basswood American beech American chestnut American elm American hornbeam Bear oak (scrub oak) Bigtooth aspen Biternut hickory Black cherry Black hickory Black locust Black oak Black walnut Black willow Blackgum Blackjack oak Boxelder Buternut Cedar elm Chestnut oak Chinkapin oak Chokecherry Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Northern red oak Osage-orange Pawpaw Pignut hickory Pin cherry Pitch pine Post oak 30 52 9 15 10 14 8 3 168 0 129 167 42 1 148 6 20 16 0 490 14 4 3 11 1 27 33 31 38 69 155 2 24 0 1 80 249 2 4 80 7 43 0 DiSTRiB Current Model iV Reliability 31 63 12 43 29 7 13 0 175 1 114 148 51 1 160 2 26 4 0 388 8 1 3 23 0 56 44 40 39 104 162 8 18 1 7 92 222 1 8 98 1 39 3 Medium High Medium Medium Medium Low High Low High High Low High Medium Low High Medium Medium Low Low High Medium Low Medium High Low High Medium Medium Medium High High Medium Medium Low High High High Medium Low High Medium High High 2010-2039 PCM GFDL B1 A1Fi 27 61 8 26 14 8 8 2 136 2 120 156 48 0 153 6 20 11 0 423 9 3 9 14 0 41 37 50 41 82 160 7 19 1 4 89 213 2 4 91 2 40 22 24 54 8 52 18 8 2 13 107 38 105 211 47 10 150 42 22 3 0 348 27 0 50 18 1 43 41 127 55 66 174 13 26 2 5 95 187 4 5 90 1 37 176 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 21 58 8 26 12 7 2 2 101 5 111 156 49 0 158 10 17 4 0 406 11 2 15 16 0 38 34 61 41 72 171 6 16 1 8 87 186 2 3 90 1 38 39 25 67 8 28 12 7 1 4 94 6 104 155 48 0 159 10 18 3 0 377 13 1 20 14 0 38 36 70 41 67 172 6 17 0 4 86 177 2 3 94 1 36 46 20 63 8 76 31 7 0 35 72 102 88 289 48 30 133 121 25 1 18 238 36 0 79 18 3 40 62 181 63 47 164 26 42 14 20 100 148 8 4 78 1 34 378 21 54 8 77 35 8 0 37 73 124 67 278 40 30 126 169 28 0 68 218 35 0 103 17 11 39 67 203 56 47 127 40 52 23 52 107 131 15 2 83 1 33 505 0.87 0.97 0.67 0.61 0.48 1.14 0.62 New 0.78 2.00 1.05 1.05 0.94 0.00 0.96 3.00 0.77 2.75 NA 1.09 1.13 3.00 3.00 0.61 NA 0.73 0.84 1.25 1.05 0.79 0.99 0.88 1.06 1.00 0.57 0.97 0.96 2.00 0.50 0.93 2.00 1.03 7.33 0.77 0.86 0.67 1.21 0.62 1.14 0.15 New 0.61 38.00 0.92 1.43 0.92 10.00 0.94 21.00 0.85 0.75 NA 0.90 3.38 0.00 16.67 0.78 New 0.77 0.93 3.18 1.41 0.64 1.07 1.63 1.44 2.00 0.71 1.03 0.84 4.00 0.63 0.92 1.00 0.95 58.67 0.68 0.92 0.67 0.61 0.41 1.00 0.15 New 0.58 5.00 0.97 1.05 0.96 0.00 0.99 5.00 0.65 1.00 NA 1.05 1.38 2.00 5.00 0.70 NA 0.68 0.77 1.53 1.05 0.69 1.06 0.75 0.89 1.00 1.14 0.95 0.84 2.00 0.38 0.92 1.00 0.97 13.00 0.65 1.00 0.67 1.77 1.07 1.00 0.00 New 0.41 102.00 0.77 1.95 0.94 30.00 0.83 60.50 0.96 0.25 New 0.61 4.50 0.00 26.33 0.78 New 0.71 1.41 4.53 1.62 0.45 1.01 3.25 2.33 14.00 2.86 1.09 0.67 8.00 0.50 0.80 1.00 0.87 126.00 0.81 1.06 0.67 0.65 0.41 1.00 0.08 4.00 0.54 6.00 0.91 1.05 0.94 0.00 0.99 5.00 0.69 0.75 0.00 0.97 1.63 1.00 6.67 0.61 0.00 0.68 0.82 1.75 1.05 0.64 1.06 0.75 0.94 0.00 0.57 0.94 0.80 2.00 0.38 0.96 1.00 0.92 15.33 0.68 0.86 0.67 1.79 1.21 1.14 0.00 37.00 0.42 124.00 0.59 1.88 0.78 30.00 0.79 84.50 1.08 0.00 68.00 0.56 4.38 0.00 34.33 0.74 11.00 0.70 1.52 5.08 1.44 0.45 0.78 5.00 2.89 23.00 7.43 1.16 0.59 15.00 0.25 0.85 1.00 0.85 168.33 Change Class 2070-2099 PCM GFDL B1 A1Fi Decrease No Change Decrease Decrease Large Decrease No Change Large Decrease Increase Decrease New Habitat No Change No Change No Change NA No Change Increase Decrease Decrease NA No change Increase No Change Increase Decrease NA Decrease Decrease Increase No Change Decrease No Change Decrease No Change NA Decrease No Change Decrease No Change Large Decrease No Change No Change No Change New Habitat Decrease No Change Decrease Increase Increase No Change Exirpated Increase Large Decrease New Habitat Decrease Increase Decrease Increase Decrease Increase No Change Exirpated New Habitat Decrease Large Increase Exirpated Increase Decrease Increase Decrease Increase Large Increase Increase Decrease Decrease Large Increase Large Increase New Habitat Large Increase Increase Decrease Increase Large Decrease Decrease No Change Decrease New Habitat (coninued on next page) APPENDiX 4 2 Table 33.—Complete DiSTRiB model results for the 73 tree speciesa in Secion M221A Table 33 (coninued). Common Name FiA iV Quaking aspen Red maple Red mulberry Red spruce Sassafras Scarlet oak Serviceberry Shagbark hickory Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern red oak Striped maple Sugar maple Sugarberry Sweet birch Sweetgum Sycamore Table Mountain pine Tulip tree Virginia pine Water locust Water oak White ash White oak Winged elm Yellow birch Yellow buckeye 6 293 2 3 76 94 77 8 7 0 24 55 5 2 44 146 0 90 0 24 21 69 133 0 0 102 254 0 8 1 DiSTRiB Current Model iV Reliability 4 357 3 3 99 114 63 14 6 0 13 51 22 1 45 173 0 98 4 23 20 119 109 0 0 146 254 0 7 5 High High Low High High High Medium Medium High Low Medium Medium High High High High Medium High High Medium Medium High High Low High High High High High Medium 2010-2039 PCM GFDL B1 A1Fi 3 333 2 2 91 110 69 12 7 0 14 48 24 3 36 151 0 83 3 22 20 97 129 0 0 108 265 4 7 1 1 272 9 0 109 107 55 27 42 1 23 51 15 15 27 166 12 62 10 30 25 102 102 0 0 114 304 44 4 3 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 3 328 1 1 97 110 58 15 10 0 12 38 35 6 26 147 3 74 6 20 20 120 129 0 0 94 260 9 6 2 3 324 1 1 107 108 52 17 10 0 14 37 32 5 25 154 4 75 9 24 18 134 116 0 0 97 254 14 6 1 0 179 32 0 97 83 52 50 126 19 46 45 22 35 20 121 52 47 13 39 27 79 86 2 7 88 300 146 4 4 0 136 40 0 90 42 53 60 197 34 58 41 23 62 21 59 78 46 37 40 30 64 85 9 46 80 244 242 4 4 0.75 0.93 0.67 0.67 0.92 0.97 1.10 0.86 1.17 NA 1.08 0.94 1.09 3.00 0.80 0.87 NA 0.85 0.75 0.96 1.00 0.82 1.18 NA NA 0.74 1.04 New 1.00 0.20 0.25 0.76 3.00 0.00 1.10 0.94 0.87 1.93 7.00 New 1.77 1.00 0.68 15.00 0.60 0.96 New 0.63 2.50 1.30 1.25 0.86 0.94 NA NA 0.78 1.20 New 0.57 0.60 0.75 0.00 0.92 0.50 0.33 10.67 0.33 0.00 0.98 0.98 0.97 0.73 0.92 0.83 1.07 3.57 1.67 21.00 NA New 0.92 3.54 0.75 0.88 1.59 1.00 6.00 35.00 0.58 0.44 0.85 0.70 New New 0.76 0.48 1.50 3.25 0.87 1.70 1.00 1.35 1.01 0.66 1.18 0.79 NA New NA New 0.64 0.60 1.02 1.18 New New 0.86 0.57 0.40 0.80 0.75 0.91 0.33 0.33 1.08 0.95 0.83 1.21 1.67 0.00 1.08 0.73 1.46 5.00 0.56 0.89 4.00 0.77 2.25 1.04 0.90 1.13 1.06 0.00 0.00 0.66 1.00 14.00 0.86 0.20 0.00 0.38 13.33 0.00 0.91 0.37 0.84 4.29 32.83 34.00 4.46 0.80 1.05 62.00 0.47 0.34 78.00 0.47 9.25 1.74 1.50 0.54 0.78 9.00 46.00 0.55 0.96 242.00 0.57 0.80 Change Class 2070-2099 PCM GFDL B1 A1Fi Decrease No Change Decrease Decrease No Change No Change Decrease Increase Increase NA No Change Decrease Increase Increase Decrease No Change New Habitat Decrease New Habitat No Change No Change No Change No Change NA NA Decrease No Change New Habitat No Change Large Decrease Exirpated Large Decrease Increase Exirpated No Change Large Decrease Decrease Large Increase Large Increase New Habitat Large Increase Decrease No Change Increase Decrease Large Decrease New Habitat Decrease New Habitat Increase Increase Decrease Decrease New Habitat New Habitat Decrease No Change New Habitat Decrease Decrease Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 27 Common Name FiA iV American basswood American beech American chestnut American elm American holly American hornbeam Balsam ir Bear oak (scrub oak) Bigtooth aspen Biternut hickory Black cherry Black hickory Black locust Black oak Black walnut Black willow Blackgum Blackjack oak Boxelder Buternut Cedar elm Chestnut oak Chinkapin oak Chokecherry Common persimmon Cucumbertree Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Northern red oak Ohio buckeye Pawpaw Pignut hickory Pin cherry Pitch pine 69 383 39 14 5 35 3 2 17 5 323 0 161 100 28 2 108 0 10 12 0 339 2 6 0 60 110 58 8 10 86 132 13 3 9 0 83 312 6 5 83 11 24 DiSTRiB Current Model iV Reliability 60 362 32 27 1 45 7 8 22 2 339 2 162 115 31 1 160 0 14 3 0 384 0 10 0 54 152 62 11 11 121 166 15 1 1 4 92 303 4 2 94 8 36 Medium High Medium Medium High Medium High Low High Low High High Low High Medium Low High Medium Medium Low Low High Medium Low Medium High High Medium Medium Medium High High Medium Medium Low High High High Low Low High Medium High 2010-2039 PCM GFDL B1 A1Fi 58 342 37 24 3 33 7 2 16 2 288 1 165 115 30 1 153 0 5 10 0 378 2 5 3 54 129 62 10 12 102 172 8 1 4 0 87 297 4 3 83 6 29 53 231 37 29 2 38 1 3 17 6 275 2 161 161 50 1 180 3 9 6 0 405 9 4 15 55 129 61 82 31 120 227 14 1 6 1 101 267 4 3 99 0 42 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 53 340 34 24 2 34 7 1 10 3 268 1 161 121 35 1 164 0 8 9 0 375 2 5 5 54 131 58 13 16 107 190 8 1 3 4 91 283 4 3 88 3 31 55 306 35 22 2 34 3 3 13 3 269 1 154 124 39 1 171 0 7 7 0 378 1 7 8 57 133 58 18 19 103 197 8 1 3 5 92 274 5 3 90 2 32 41 202 37 56 1 50 1 5 4 33 159 63 145 243 57 4 211 60 10 3 0 340 31 0 85 50 106 74 239 59 128 278 20 13 14 23 137 249 2 0 109 0 44 30 168 36 78 1 53 1 4 0 53 110 143 121 305 63 12 221 133 16 0 17 274 46 0 147 49 93 92 290 63 110 271 32 30 23 80 157 234 0 0 104 0 47 0.97 0.95 1.16 0.89 3.00 0.73 1.00 0.25 0.73 1.00 0.85 0.50 1.02 1.00 0.97 1.00 0.96 NA 0.36 3.33 NA 0.98 New 0.50 New 1.00 0.85 1.00 0.91 1.09 0.84 1.04 0.53 1.00 4.00 0.00 0.95 0.98 1.00 1.50 0.88 0.75 0.81 0.88 0.64 1.16 1.07 2.00 0.84 0.14 0.38 0.77 3.00 0.81 1.00 0.99 1.40 1.61 1.00 1.13 New 0.64 2.00 NA 1.06 New 0.40 New 1.02 0.85 0.98 7.46 2.82 0.99 1.37 0.93 1.00 6.00 0.25 1.10 0.88 1.00 1.50 1.05 0.00 1.17 0.88 0.94 1.06 0.89 2.00 0.76 1.00 0.13 0.46 1.50 0.79 0.50 0.99 1.05 1.13 1.00 1.03 NA 0.57 3.00 NA 0.98 New 0.50 New 1.00 0.86 0.94 1.18 1.46 0.88 1.15 0.53 1.00 3.00 1.00 0.99 0.93 1.00 1.50 0.94 0.38 0.86 0.68 0.56 1.16 2.07 1.00 1.11 0.14 0.63 0.18 16.50 0.47 31.50 0.90 2.11 1.84 4.00 1.32 New 0.71 1.00 NA 0.89 New 0.00 New 0.93 0.70 1.19 21.73 5.36 1.06 1.68 1.33 13.00 14.00 5.75 1.49 0.82 0.50 0.00 1.16 0.00 1.22 0.92 0.85 1.09 0.82 2.00 0.76 0.43 0.38 0.59 1.50 0.79 0.50 0.95 1.08 1.26 1.00 1.07 NA 0.50 2.33 NA 0.98 New 0.70 New 1.06 0.88 0.94 1.64 1.73 0.85 1.19 0.53 1.00 3.00 1.25 1.00 0.90 1.25 1.50 0.96 0.25 0.89 0.50 0.46 1.13 2.89 1.00 1.18 0.14 0.50 0.00 26.50 0.32 71.50 0.75 2.65 2.03 12.00 1.38 New 1.14 0.00 New 0.71 New 0.00 New 0.91 0.61 1.48 26.36 5.73 0.91 1.63 2.13 30.00 23.00 20.00 1.71 0.77 0.00 0.00 1.11 0.00 1.31 Change Class 2070-2099 PCM GFDL B1 A1Fi No Change Decrease No Change Decrease No Change Decrease Large Decrease Large Decrease Decrease No Change Decrease Decrease No Change No Change Increase No Change No Change NA Decrease No Change NA No Change No Change Decrease New Habitat No Change No Change No Change Increase Increase No Change Increase Decrease No Change No Change No Change No Change No Change No Change No Change No Change Large Decrease No Change Decrease Decrease No Change Large Increase No Change Increase Large Decrease Decrease Exirpated Increase Large Decrease Increase Decrease Large Increase Increase Increase Increase New Habitat No Change Exirpated New Habitat Decrease Increase Exirpated New Habitat No Change Decrease Increase Large Increase Large Increase No Change Increase Large Increase Increase Increase Increase Increase Decrease Exirpated Exirpated No Change Exirpated Increase (coninued on next page) APPENDiX 4 28 Table 34.—Complete DiSTRiB model results for the 74 tree speciesa in Secion M221B Table 34 (coninued). Common Name FiA iV Post oak Quaking aspen Red maple Red mulberry Red pine Red spruce Sassafras Scarlet oak Serviceberry Shagbark hickory Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern red oak Striped maple Sugar maple Sugarberry Sweet birch Sweetgum Sycamore Table Mountain pine Tulip tree Virginia pine Water oak White ash White oak Winged elm Yellow birch Yellow buckeye 3 2 645 0 15 64 121 103 101 17 4 0 1 25 64 2 154 483 1 193 0 15 5 204 38 0 125 221 0 102 23 DiSTRiB Current Model iV Reliability 0 17 686 0 9 35 137 126 88 12 6 0 1 26 87 0 138 452 0 196 0 12 15 232 63 0 141 214 0 99 21 High High High Low Medium High High High Medium Medium High Low Medium Medium High High High High Medium High High Medium Medium High High High High High High High Medium 2010-2039 PCM GFDL B1 A1Fi 4 7 663 0 10 38 128 132 91 13 8 0 0 25 84 0 129 447 0 200 0 16 10 246 60 0 123 226 0 80 23 27 4 640 3 7 20 156 140 82 20 21 0 1 42 96 4 92 356 0 183 3 28 11 278 95 0 143 280 7 41 25 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 6 5 657 0 10 32 131 138 89 13 9 0 1 26 102 1 119 433 0 196 0 18 12 275 68 0 109 228 0 80 25 8 4 660 1 8 33 132 137 86 15 10 0 0 29 107 0 114 415 0 196 1 18 11 297 71 0 109 239 0 79 25 211 1 497 29 0 15 154 155 71 51 120 2 7 65 81 32 60 318 15 139 42 45 16 270 128 3 129 398 93 23 28 437 0 348 44 0 14 140 126 71 79 264 19 14 72 72 79 50 230 34 104 85 51 21 195 126 17 113 424 207 20 27 New 0.41 0.97 NA 1.11 1.09 0.93 1.05 1.03 1.08 1.33 NA 0.00 0.96 0.97 NA 0.94 0.99 NA 1.02 NA 1.33 0.67 1.06 0.95 NA 0.87 1.06 NA 0.81 1.10 New 0.24 0.93 New 0.78 0.57 1.14 1.11 0.93 1.67 3.50 NA 1.00 1.62 1.10 New 0.67 0.79 NA 0.93 New 2.33 0.73 1.20 1.51 NA 1.01 1.31 New 0.41 1.19 New New 0.29 0.06 0.96 0.72 NA New 1.11 0.00 0.91 0.43 0.96 1.12 1.10 1.23 1.01 0.81 1.08 4.25 1.50 20.00 NA New 1.00 7.00 1.00 2.50 1.17 0.93 New New 0.86 0.44 0.96 0.70 NA New 1.00 0.71 NA New 1.50 3.75 0.80 1.07 1.19 1.16 1.08 2.03 NA New 0.77 0.92 1.07 1.86 NA New 0.81 0.23 1.19 1.33 New New 0.24 0.00 0.96 0.51 New New 0.89 0.00 0.94 0.40 0.96 1.02 1.09 1.00 0.98 0.81 1.25 6.58 1.67 44.00 NA New 0.00 14.00 1.12 2.77 1.23 0.83 NA New 0.83 0.36 0.92 0.51 NA New 1.00 0.53 New New 1.50 4.25 0.73 1.40 1.28 0.84 1.13 2.00 NA New 0.77 0.80 1.12 1.98 NA New 0.80 0.20 1.19 1.29 Change Class 2070-2099 PCM GFDL B1 A1Fi Increase Large Decrease No Change New Habitat No Change No Change No Change No Change No Change Increase Increase NA Exirpated No Change Increase Exirpated Decrease No Change Exirpated No Change New Habitat Increase Decrease Increase No Change NA Decrease No Change NA Decrease Increase Increase Exirpated Decrease New Habitat Exirpated Large Decrease No Change No Change Decrease Large Increase Large Increase New Habitat Increase Large Increase Decrease Increase Large Decrease Decrease Increase Decrease New Habitat Large Increase Increase Decrease Increase New Habitat Decrease Increase New Habitat Large Decrease Increase Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 29 Common Name FiA iV American basswood American beech American chestnut American elm American hornbeam Biternut hickory Black cherry Black hickory Black locust Black oak Black walnut Blackgum Blackjack oak Buternut Cedar elm Chestnut oak Chinkapin oak Common persimmon Cucumbertree Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Loblolly pine Mockernut hickory Mountain maple Northern red oak Pawpaw Pignut hickory Pitch pine Post oak Red maple Red mulberry River birch Sassafras Scarlet oak Serviceberry Shagbark hickory Shortleaf pine 81 227 19 20 39 5 56 0 83 118 22 103 0 11 0 238 5 6 68 72 40 2 29 18 207 7 0 0 84 12 154 42 85 13 2 417 5 5 210 88 30 10 7 DiSTRiB Current Model iV Reliability 52 222 11 16 40 0 91 0 94 118 27 126 1 1 0 261 2 2 55 62 37 4 32 35 193 6 0 2 103 3 156 23 104 18 6 448 2 1 177 83 42 5 11 Medium High Medium Medium Medium Low High High Low High Medium High Medium Low Low High Medium Medium High High Medium Medium Medium High High Medium Medium High High High High Low High High High High Low Low High High Medium Medium High 2010-2039 PCM GFDL B1 A1Fi 66 215 17 19 37 5 63 0 90 107 27 125 1 7 0 243 2 6 58 72 38 6 34 34 204 6 0 2 91 8 149 26 89 12 6 427 3 3 180 82 35 9 15 43 188 15 22 41 11 48 29 93 157 35 135 21 0 0 231 14 26 53 60 41 45 47 42 220 7 2 14 101 10 139 20 91 17 78 360 6 2 159 96 38 24 67 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 60 213 16 18 37 4 54 0 93 106 27 124 0 5 0 248 2 3 55 71 37 8 38 43 209 6 0 7 93 8 139 26 88 11 9 425 3 2 166 85 35 9 19 59 214 14 18 38 5 52 0 96 110 28 128 0 5 0 241 3 5 56 70 38 12 38 47 202 7 1 8 94 8 139 21 87 10 14 427 4 2 169 82 36 10 21 28 137 10 32 42 35 43 116 75 219 50 140 109 0 21 161 32 73 47 49 60 101 52 30 209 12 11 66 111 10 166 9 69 13 305 236 17 2 106 68 38 62 250 31 118 10 46 41 37 59 151 70 235 67 134 175 0 72 156 37 95 44 48 75 141 69 30 154 25 28 139 137 10 155 8 71 20 507 168 35 3 93 38 38 82 342 1.27 0.97 1.55 1.19 0.93 New 0.69 NA 0.96 0.91 1.00 0.99 1.00 7.00 NA 0.93 1.00 3.00 1.06 1.16 1.03 1.50 1.06 0.97 1.06 1.00 NA 1.00 0.88 2.67 0.96 1.13 0.86 0.67 1.00 0.95 1.50 3.00 1.02 0.99 0.83 1.80 1.36 0.83 0.85 1.36 1.38 1.03 New 0.53 New 0.99 1.33 1.30 1.07 21.00 0.00 NA 0.89 7.00 13.00 0.96 0.97 1.11 11.25 1.47 1.20 1.14 1.17 New 7.00 0.98 3.33 0.89 0.87 0.88 0.94 13.00 0.80 3.00 2.00 0.90 1.16 0.91 4.80 6.09 1.15 0.96 1.46 1.13 0.93 New 0.59 NA 0.99 0.90 1.00 0.98 0.00 5.00 NA 0.95 1.00 1.50 1.00 1.15 1.00 2.00 1.19 1.23 1.08 1.00 NA 3.50 0.90 2.67 0.89 1.13 0.85 0.61 1.50 0.95 1.50 2.00 0.94 1.02 0.83 1.80 1.73 0.54 0.62 0.91 2.00 1.05 New 0.47 New 0.80 1.86 1.85 1.11 109.00 0.00 New 0.62 16.00 36.50 0.86 0.79 1.62 25.25 1.63 0.86 1.08 2.00 New 33.00 1.08 3.33 1.06 0.39 0.66 0.72 50.83 0.53 8.50 2.00 0.60 0.82 0.91 12.40 22.73 1.14 0.96 1.27 1.13 0.95 New 0.57 New 1.02 0.93 1.04 1.02 0.00 5.00 New 0.92 1.50 2.50 1.02 1.13 1.03 3.00 1.19 1.34 1.05 1.17 New 4.00 0.91 2.67 0.89 0.91 0.84 0.56 2.33 0.95 2.00 2.00 0.96 0.99 0.86 2.00 1.91 0.60 0.53 0.91 2.88 1.03 New 0.65 New 0.75 1.99 2.48 1.06 175.00 0.00 New 0.60 18.50 47.50 0.80 0.77 2.03 35.25 2.16 0.86 0.80 4.17 New 69.50 1.33 3.33 0.99 0.35 0.68 1.11 84.50 0.38 17.50 3.00 0.53 0.46 0.91 16.40 31.09 Change Class 2070-2099 PCM GFDL B1 A1Fi No Change No Change Increase No Change No Change No Change Decrease NA No Change No Change No Change No Change NA Increase NA No Change No Change No Change No Change No Change No Change Large Increase Increase Increase No Change Increase NA New Habitat No Change No Change No Change No Change Decrease Decrease No Change No Change No Change Increase No Change No Change No Change Increase Increase Decrease Decrease No Change Large Increase No Change Large Increase Decrease New Habitat Decrease Increase Large Increase No Change New Habitat Exirpated New Habitat Decrease Increase Increase Decrease Decrease Increase Large Increase Large Increase No Change Decrease Large Increase New Habitat New Habitat Increase No Change No Change Large Decrease Decrease No Change Increase Large Decrease Increase Large Increase Decrease Decrease No Change Large Increase Large Increase (coninued on next page) APPENDiX 4 270 Table 35.—Complete DiSTRiB model results for the 62 tree speciesa in Secion M221C Table 35 (coninued). Common Name FiA iV Shumard oak Slippery elm Sourwood Southern red oak Striped maple Sugar maple Sugarberry Sweet birch Sweetgum Sycamore Tulip tree Virginia pine Water locust Water oak White ash White oak Winged elm Yellow birch Yellow buckeye 1 35 94 2 38 355 0 101 3 17 361 14 0 0 70 145 0 26 31 DiSTRiB Current Model iV Reliability 0 24 100 1 32 288 0 100 2 15 323 42 0 0 68 160 1 17 29 Low Medium High High High High Medium High High Medium High High Low High High High High High Medium 2010-2039 PCM GFDL B1 A1Fi 0 34 102 2 31 331 0 99 8 16 356 40 0 0 64 158 1 19 33 0 31 113 22 18 249 1 75 28 24 340 79 0 0 66 205 35 7 34 Modeled iV 2040-2069 PCM GFDL B1 A1Fi DiSTRiB results Future : Current Suitable habitat 2070-2099 2010-2039 2040-2069 2070-2099 PCM GFDL PCM GFDL PCM GFDL PCM GFDL B1 A1Fi B1 A1Fi B1 A1Fi B1 A1Fi 0 31 119 6 27 313 0 95 11 16 357 60 0 0 61 163 5 16 33 0 34 119 9 24 316 0 89 16 18 357 68 0 0 61 176 7 16 34 14 35 98 65 18 129 27 48 69 30 186 93 1 21 66 262 153 6 33 35 37 71 93 19 53 43 44 82 39 95 87 9 57 75 230 235 5 33 NA 1.42 1.02 2.00 0.97 1.15 NA 0.99 4.00 1.07 1.10 0.95 NA NA 0.94 0.99 1.00 1.12 1.14 NA 1.29 1.13 22.00 0.56 0.87 New 0.75 14.00 1.60 1.05 1.88 NA NA 0.97 1.28 35.00 0.41 1.17 NA New 1.29 1.46 1.19 0.98 6.00 65.00 0.84 0.56 1.09 0.45 NA New 0.95 0.48 5.50 34.50 1.07 2.00 1.11 0.58 1.43 2.21 NA New NA New 0.90 0.97 1.02 1.64 5.00 153.00 0.94 0.35 1.14 1.14 New New 1.42 1.54 1.19 0.71 9.00 93.00 0.75 0.59 1.10 0.18 New New 0.89 0.44 8.00 41.00 1.20 2.60 1.11 0.29 1.62 2.07 New New New New 0.90 1.10 1.10 1.44 7.00 235.00 0.94 0.29 1.17 1.14 Change Class 2070-2099 PCM GFDL B1 A1Fi NA Increase Increase Increase Decrease No Change NA No Change Increase Increase No Change Increase NA NA No Change No Change New Habitat No Change Increase Increase Increase Decrease Increase Decrease Large Decrease New Habitat Large Decrease Increase Large Increase Large Decrease Large Increase New Habitat New Habitat No Change Increase New Habitat Large Decrease No Change Current importance values (Current IV) are based on results from the DISTRIB model. Early-, mid-, and late-century importance values are average values for the indicated years. Change classes are provided for the end of century (2070 through 2099) period. Explanaions for the change classes are described in the text. Future:Current Suitable Habitat is a raio of projected importance value to current importance value. a APPENDiX 4 271 APPENDiX 4 Table 36.—Comparison of change classes for two climate scenarios from the DiSTRiB model results for all tree species in six ecological secions of the assessment areaa 221F Common Name American basswood American beech American chestnut American elm American holly American hornbeam Balsam ir Bear oak (scrub oak) PCM B1 Ecological secion within the assessment area boundaries 221E OH 221E WV GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi Decrease Large Decrease No Change Large Decrease Decrease Large Decrease No Change Large Decrease No Change - - - Large Decrease Decrease Large Decrease - No Change Exirpated No Change Decrease Decrease Decrease Decrease No Change - - - - - - No Change No Change No Change Increase Decrease No Change - - - - - - - - - - - - Decrease Exirpated Large Decrease Exirpated Large Decrease Exirpated Biternut hickory Large Increase Large Increase Increase Large Increase Increase Large Increase Black ash Bigtooth aspen Large Decrease Large Decrease - - - - Black cherry Decrease Large Decrease Decrease Large Decrease Large Decrease Large Decrease Black hickory NA New New New New New Black locust Increase No Change Decrease Decrease Decrease Decrease Black maple No Change Large Decrease No Change Large Decrease Decrease Decrease Black oak Increase Large Increase Increase Large Increase Decrease Increase Black walnut Increase Increase No Change Decrease Increase Decrease Black willow No Change Increase Large Decrease Large Increase NA Increase Increase No Change Increase No Change Decrease No Change New New Increase Increase Increase Increase Blackgum Blackjack oak Boxelder No Change Increase Decrease Increase Decrease No Change Bur oak Increase Large Increase Large Decrease Large Increase - - - - - - No Change Exirpated Buternut NA New NA New NA New Chestnut oak New Large Decrease Increase Decrease Decrease Decrease Chinkapin oak New New Large Increase Large Increase No Change Increase Cedar elm Chokecherry Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redbud Eastern redcedar - - - - - - New New Increase Large Increase Increase Large Increase Large Increase No Change Large Increase Increase Increase No Change Decrease Large Increase Decrease Large Increase - - Large Decrease Large Decrease Large Decrease Large Decrease Decrease Large Decrease Decrease Decrease No Change Increase Decrease Increase New New Increase Increase Decrease No Change New New Large Increase Large Increase Large Increase Large Increase Eastern white pine No Change Exirpated Decrease Exirpated Increase Exirpated Flowering dogwood Increase No Change Increase Decrease Decrease Large Decrease Green ash Increase Large Increase Increase Large Increase Large Increase Large Increase Hackberry Large Increase Large Increase Increase Large Increase Increase Large Increase Honeylocust Large Increase Large Increase No Change Large Increase No Change Large Increase (coninued on next page) 272 APPENDiX 4 Table 36 (coninued). 221F Common Name Loblolly pine Mockernut hickory Mountain maple PCM B1 Ecological secion within the assessment area boundaries 221E OH 221E WV GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi - - NA New Increase Increase No Change Increase No Change Increase Decrease Increase - - - - - - Northern catalpa - - No Change No Change - - Northern pin oak No Change Decrease - - - - Northern red oak No Change Decrease No Change Decrease Decrease No Change Decrease No Change - - - - Northern white-cedar Ohio buckeye Large Increase Decrease No Change Large Decrease Decrease Exirpated Osage-orange Increase Large Increase No Change Large Increase No Change Increase - - No Change Large Decrease Decrease Large Decrease No Change Increase Increase No Change Decrease Decrease Decrease Large Decrease - - Exirpated Exirpated No Change No Change No Change Large Increase New New Pawpaw Pignut hickory Pin cherry Pin oak Pitch pine - - Increase No Change Decrease Large Decrease Post oak New New Large Increase Large Increase Large Increase Large Increase Large Decrease Large Decrease Exirpated Exirpated - - Decrease Large Decrease No Change Large Decrease Decrease Large Decrease New New NA New No Change Large Increase Decrease Large Decrease Large Decrease Large Decrease - - Quaking aspen Red maple Red mulberry Red pine Red spruce - - - - - - River birch - - No Change Increase - - Rock elm NA New - - Large Increase Large Increase Sassafras Increase Increase No Change Decrease Decrease Decrease Large Increase Large Increase Increase Decrease Decrease Large Decrease No Change Decrease No Change No Change Decrease No Change Increase Increase Increase Increase Increase Large Increase Scarlet oak Serviceberry Shagbark hickory No Change Increase No Change Increase - - Shortleaf pine Shingle oak NA New Large Increase Large Increase Large Increase Large Increase Shumard oak NA New NA NA NA New Silver maple No Change Increase Decrease Decrease Large Decrease Large Increase Slippery elm Increase Decrease No Change No Change Decrease Large Decrease - - Increase Increase Increase Large Decrease NA New New New Increase Increase - - - - - - Sugar maple No Change Large Decrease No Change No Change Decrease Large Decrease Sugarberry NA New New New New New Swamp white oak No Change Large Decrease - - - - Sweet birch No Change Exirpated Decrease Decrease Decrease Large Decrease Sweetgum New New Increase Increase Increase Increase Sycamore Increase Increase No Change No Change Decrease Sourwood Southern red oak Striped maple No Change (coninued on next page) 273 APPENDiX 4 Table 36 (coninued). 221F Common Name Table Mountain pine PCM B1 Ecological secion within the assessment area boundaries 221E OH 221E WV GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi - - - - No Change No Change Increase Decrease Increase Increase Decrease Large Decrease Virginia pine - - Increase Increase Increase Large Decrease Water oak - - NA NA NA New Water locust - - NA NA NA New White ash No Change Large Decrease Decrease Decrease Decrease Large Decrease White oak No Change Increase Increase Increase Decrease No Change Willow oak - - - - NA New Winged elm NA New New New Increase Increase Yellow birch Large Decrease Exirpated - - - - - - No Change No Change Decrease Tulip tree Yellow buckeye Change classes are provided for the end-of-century (2070 through 2099) period. Explanaions for the change classes are described in the text. Blue ash, southern magnolia, and tamarack were present only at the regional level and do not appear here. a 274 No Change (coninued on next page) APPENDiX 4 Table 36 (coninued).—Comparison of change classes for two climate scenarios from the DISTRIB model results for all tree species in six ecological secions of the assessment areaa Common Name Ecological secion within the assessment area boundaries M221C M221B M221A PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi American basswood No Change Decrease No Change Decrease Decrease Decrease American beech No Change Decrease Decrease Decrease No Change No Change Increase No Change No Change No Change Decrease Decrease American elm No Change Large Increase Decrease Large Increase Decrease Increase American holly - - No Change No Change - - American chestnut No Change No Change Decrease Increase Large Decrease Increase Balsam ir American hornbeam - - Large Decrease Large Decrease - - Bear oak (scrub oak) - - Large Decrease Decrease No Change No Change Bigtooth aspen Biternut hickory Black ash - - Decrease Exirpated Large Decrease Exirpated No Change Large Increase No Change Increase Increase Increase - - - - - - Black cherry Decrease Decrease Decrease Large Decrease Decrease Large Decrease Black hickory NA New Decrease Increase New New Black locust No Change Decrease No Change Decrease No Change Decrease Black maple - - - - - - Black oak No Change Increase No Change Large Increase No Change Increase Black walnut No Change Large Increase Increase Increase No Change Decrease Black willow - - No Change Increase NA Increase No Change No Change No Change Increase No Change Decrease Blackgum Blackjack oak NA New NA New Increase Increase Boxelder - - Decrease No Change Decrease No Change Bur oak - - - - - - Buternut Increase Exirpated No Change Exirpated Decrease Exirpated Cedar elm NA New NA New NA New Chestnut oak No Change Decrease No Change Decrease No Change Decrease Chinkapin oak No Change Increase No Change Increase Increase Large Increase Chokecherry - - Decrease Exirpated No Change Exirpated Common persimmon No Change Increase New New Increase Increase Cucumbertree No Change Decrease No Change No Change Decrease Decrease - - - - NA Increase Eastern hemlock No Change Decrease No Change Decrease Decrease Decrease Eastern hophornbeam No Change Increase No Change Increase Decrease Increase Increase Large Increase Increase Large Increase No Change Increase Large Increase Large Increase Increase Large Increase Increase Large Increase Eastern white pine Increase No Change No Change No Change Decrease Decrease Flowering dogwood Eastern cotonwood Eastern redbud Eastern redcedar No Change Decrease Increase Increase No Change Decrease Green ash Increase Large Increase Decrease Large Increase Decrease Large Increase Hackberry NA New No Change Increase No Change Large Increase - - No Change Increase NA New Honeylocust (coninued on next page) 275 APPENDiX 4 Table 36 (coninued). Common Name Loblolly pine Ecological secion within the assessment area boundaries M221C M221B M221A PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi New New No Change Increase Decrease Large Increase Mockernut hickory No Change Increase No Change Increase No Change Increase Mountain maple No Change No Change - - - - Northern catalpa - - - - - - Northern pin oak - - - - - - Northern red oak No Change No Change No Change Decrease Decrease Decrease Northern white-cedar - - - - - - Ohio buckeye - - No Change Exirpated - - Osage-orange Pawpaw Pignut hickory Pin cherry - - - - No Change Increase No Change Large Decrease No Change Exirpated Large Decrease Large Decrease Decrease Decrease No Change No Change No Change Decrease - - Large Decrease Exirpated No Change No Change - - - - - - Pitch pine Decrease No Change No Change Increase No Change Decrease Post oak No Change Increase Increase Increase New New Pin oak - - Large Decrease Exirpated Decrease Exirpated Red maple No Change Large Decrease No Change Decrease No Change Large Decrease Red mulberry No Change Increase New New Decrease Increase Red pine - - No Change Exirpated - - Red spruce - - No Change Large Decrease Decrease Exirpated River birch Increase Large Increase - - - - Quaking aspen Rock elm - - - - - - Sassafras No Change Decrease No Change No Change No Change No Change Scarlet oak No Change Decrease No Change No Change No Change Large Decrease Serviceberry No Change No Change No Change Decrease Decrease Decrease Increase Large Increase Increase Large Increase Increase Large Increase - - - - - - Shortleaf pine Increase Large Increase Increase Large Increase Increase Large Increase Shumard oak NA Increase NA New NA New Silver maple - - Exirpated Increase No Change Large Increase Slippery elm Increase Increase No Change Large Increase Decrease Decrease Sourwood Increase Decrease Increase Decrease Increase No Change Southern red oak Increase Increase Exirpated Increase Increase Increase Striped maple Decrease Decrease Decrease Large Decrease Decrease Decrease Sugar maple No Change Large Decrease No Change Decrease No Change Large Decrease Sugarberry NA New Exirpated Increase New New - - - - - - Sweet birch No Change Large Decrease No Change Decrease Decrease Decrease Sweetgum Increase Increase New New New New Sycamore Increase Large Increase Increase Large Increase No Change Increase Shagbark hickory Shingle oak Swamp white oak (coninued on next page) 27 APPENDiX 4 Table 36 (coninued). Common Name Table Mountain pine Tulip tree Virginia pine Ecological secion within the assessment area boundaries M221C M221B M221A PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi PCM B1 GFDL A1Fi - - Decrease Increase No Change Increase No Change Large Decrease Increase Decrease No Change Decrease Increase Large Increase No Change Increase No Change Decrease Water oak NA New NA New NA New Water locust NA New - - NA New White ash No Change No Change Decrease Decrease Decrease Decrease White oak No Change Increase No Change Increase No Change No Change Willow oak - - - - - - Winged elm New New NA New New New Yellow birch No Change Large Decrease Decrease Large Decrease No Change Decrease Increase No Change Increase Increase Large Decrease Decrease Yellow buckeye Change classes are provided for the end-of-century (2070 through 2099) period. Explanaions for the change classes are described in the text. Blue ash, southern magnolia, and tamarack were present only at the regional level and do not appear here. a 277 APPENDiX 4 Table 37.—Modifying factor and adaptability informaion for the 93 tree species in the assessment area that were modeled by using DiSTRiB Common Name American basswood American beech American chestnut American elm American holly American hornbeam Balsam ir Bear oak (scrub oak) Bigtooth aspen Biternut hickory Black ash Black cherry Black hickory Black locust Black maple Black oak Black walnut Black willow Blackgum Blackjack oak Blue ash Boxelder Bur oak Buternut Cedar elm Chestnut oak Chinkapin oak Chokecherry Common persimmon Cucumbertree Eastern cotonwood Eastern hemlock Eastern hophornbeam Eastern redcedar Eastern redbud Eastern white pine Flowering dogwood Green ash Hackberry Honeylocust Loblolly pine Mockernut hickory Mountain maple Northern catalpa Northern pin oak Northern red oak DiSTRiB Modifying Factorsa Model Reliability Posiive Traits Negaive Traits Medium High Medium Medium High Medium High Low High Low High High High Low Low High Medium Low High Medium Low Medium Medium Low Low High Medium Low Medium High Low High Medium Medium Medium High High Medium Medium Low High High High Low Medium High COL COL COL ESP COL ESP COL SES COL FRG VRE FRG DISP DRO DRO ESP COL ESP DRO ESP SES FTK INS FTK DISE FTK DISE INS FTK FTK DRO INS FTK DRO COL FTK COL DRO FTK COL INS COL DISP DRO SES FTK ESP INS FTK COL ESP COL COL INS FTK INS DISE COL DRO COL FTK DRO COL FTK DRO SES FRG VRE COL FTK INS DISP FTK COL ESP SES DISP DRO COL SES FTK DRO FTK FTK COL DRO DISE DISE SES VRE ESP FTK INS DISE SES COL COL ESP FTK SES INS COL DISE FTK COL INS DRO COL ESP SES DRO FTK COL INS DISP COL DRO ESP COL VRE ESP DRO FTK DRO FTK INS INS FTK COL FTK COL INS INP DRO COL FTK DRO FTK COL ESP COL INS Adaptability Scores DistFact BioFact Adapt Adapt Class 0.3 -1.1 0.1 -0.8 -0.1 0.6 -3.0 1.0 1.0 2.2 -1.3 -1.6 1.0 0.0 0.5 0.5 0.4 -0.3 1.5 1.6 -0.4 2.4 2.8 -1.4 -0.3 1.4 1.2 0.2 1.2 0.0 0.2 -1.3 1.7 0.6 0.9 -2.0 0.1 -0.1 1.7 1.9 -0.5 1.7 0.8 0.9 2.5 1.4 0.2 0.0 0.3 0.3 0.5 0.6 -0.4 -0.8 0.2 -0.8 -3.0 -0.3 -2.3 -0.6 0.9 0.4 -0.8 -2.1 0.8 0.2 -2.4 2.1 -0.2 -1.3 -1.2 1.3 -0.7 -0.9 1.0 -1.1 -0.8 -0.9 1.3 -1.5 0.0 0.1 1.0 -0.3 0.3 -0.5 -0.7 -0.3 1.5 -1.6 -0.6 0.1 4.6 3.6 4.5 4.0 4.5 5.1 2.7 4.6 5.1 5.6 1.7 3.0 4.1 3.8 5.2 4.9 4.0 2.8 5.9 5.6 2.7 7.4 6.4 2.3 3.3 6.1 4.8 3.8 5.8 3.6 3.9 2.7 6.4 3.9 4.9 3.3 5.0 4.0 5.7 5.5 3.4 5.4 5.9 4.2 6.0 5.4 ○ ○ ○ ○ ○ ○ ○ ○ + ○ ○ ○ ○ ○ + + + + ○ + ○ ○ + ○ ○ + ○ ○ ○ ○ ○ + + ○ + + ○ + + (coninued on next page) 278 APPENDiX 4 Table 37 (coninued). Common Name Northern white-cedar Ohio buckeye Osage-orange Pawpaw Pignut hickory Pin cherry Pin oak Pitch pine Post oak Quaking aspen Red maple Red mulberry Red pine Red spruce River birch Rock elm Sassafras Scarlet oak Serviceberry Shagbark hickory Shingle oak Shortleaf pine Shumard oak Silver maple Slippery elm Sourwood Southern magnolia Southern red oak Striped maple Sugar maple Sugarberry Swamp white oak Sweet birch Sweetgum Sycamore Table Mountain pine Tamarack (naive) Tulip tree Virginia pine Water locust Water oak White ash White oak Willow oak Winged elm Yellow birch Yellow buckeye a DiSTRiB Modifying Factorsa Model Reliability Posiive Traits Negaive Traits High Low Medium Low High Medium Medium High High High High Low Medium High Low Low High High Medium Medium Medium High Low Medium Medium High Medium High High High Medium Low High High Medium Medium High High High Low High High High Medium High High Medium COL COL ESP ESP COL ESP SES FRG FTK DRO SES FTK SES FRG ESP SES ESP ESP COL DISP COL DISP ESP COL DISP VRE ESP ESP COL SES ESP ESP DRO SES DISP SES COL COL COL ESP SES COL FTK SES COL SES COL ESP COL SES FTK SES FTK DRO INS DRO COL FTK COL INS DISE COL INS COL INS DISE COL DRO FTK FTK INS COL DISP FTK SES FTK COL DRO ESP ESP SES COL FTK INS DISE FTK DRO INS FTK COL COL INS DRO COL DRO FTK FTK DISE DRO EHS DRO FTK DISP VRE ESP FTK COL INS DISE FTK COL DRO DRO COL FTK COL INS INP COL POL SES DISP ESP SES ESP ESP SES FTK SES SES DISP COL FTK COL INS FTK COL INS DISE COL INS DISE FTK INS DISE DRO SES FTK ESP DISP Adaptability Scores DistFact BioFact Adapt Adapt Class -0.7 0.4 2.3 -0.5 0.2 0.5 -0.7 0.6 2.2 0.6 3.0 0.1 0.9 -1.3 -0.5 -0.2 0.5 -0.4 -0.4 -0.2 1.3 0.0 2.5 0.1 0.0 2.6 0.6 1.2 1.0 0.9 -0.2 1.0 -1.3 -0.4 1.3 2.6 -0.5 0.1 0.1 0.0 -0.2 -2.0 1.7 0.6 -0.6 -1.4 0.0 0.5 -1.9 0.3 -0.3 0.4 -0.7 -1.4 -1.8 -0.6 0.0 3.0 0.6 -2.4 -0.6 -0.3 -2.6 -0.6 0.7 1.0 0.4 -0.7 -1.0 -1.0 1.6 0.7 1.0 0.4 0.2 0.3 1.3 0.6 -0.3 -0.3 0.2 -0.9 -1.1 -1.2 1.3 -0.8 -0.6 -0.6 -0.5 1.0 0.0 -0.3 0.0 -2.1 4.2 3.5 6.3 3.7 4.7 4.2 2.8 3.8 5.7 4.7 8.5 4.7 3.9 2.9 3.7 2.8 4.2 4.6 4.8 4.4 4.9 3.6 5.8 5.6 4.8 6.9 4.9 5.3 5.1 5.8 4.6 4.9 3.2 4.1 4.8 5.9 3.1 5.3 3.8 3.8 3.7 2.7 6.1 4.7 3.6 3.4 3.1 ○ ○ + ○ ○ ○ ○ + ○ + ○ ○ ○ ○ ○ ○ ○ ○ ○ + + ○ + ○ + ○ + ○ ○ ○ ○ + + ○ ○ ○ + ○ ○ ○ - Modifying factor codes are described in Table 38. Adaptability scores are described in the appendix text. 279 APPENDiX 4 Table 38.—Key to modifying factor codesa Code Title Type Descripion (if posiive) Descripion (if negaive) COL Compeiion-light Biological Tolerant of shade or limited light condiions Intolerant of shade or limited light condiions DISE Disease Disturbance N/A Has a high number and/or severity of known pathogens that atack the species DISP Dispersal Biological High ability to efecively produce and distribute seeds N/A DRO Drought Biological Drought-tolerant Suscepible to drought ESP Edaphic speciicity Biological Wide range of soil requirements Narrow range of soil requirements EHS Environmental habitat speciicity Biological Wide range of suitable habitat condiions Narrow range of suitable habitat condiions FRG Fire regeneraion Disturbance Regenerates well ater ire N/A FTK Fire topkill Disturbance Resistant to ire topkill Suscepible to ire topkill INP Invasive plants Disturbance N/A Strong negaive efects of invasive plants on the species, either through compeiion for nutrients or as a pathogen INS Insect pests Disturbance N/A Has a high number and/or severity of known insects that atack the species POL Polluion Disturbance N/A Strong negaive efects of polluion on the species SES Seedling establishment Biological High ability to regenerate with seeds to maintain future populaions Low ability to regenerate with seeds to maintain future populaions VRE Vegetaive reproducion Biological Capable of vegetaive reproducion through stump sprouts or cloning N/A These codes are used to describe posiive or negaive modifying factors in Table 37. A species was given that code if informaion from the literature suggested that it had these characterisics (Mathews et al. 2011). See Mathews et al. (2011) for a more thorough descripion of these factors and how they were assessed. a 280 APPENDiX 4 6 3 biological score Length 2 High biological, low disturbance scores 1 Mid-range adaptability score: 4.25 0 -1 Lowest possible adaptability score: 0 -2 0 Highest possible adaptability score: 8.5 -3 -3 0 -2 High disturbance, low biological scores -1 0 1 disturbance score Length 2 3 6 Figure 59.—Schemaic showing how adaptability was determined for informaion for tree species modeled using the Climate Change Tree Atlas. 281 APPENDiX 4 LiNKAGES MoDEL RESuLTS Species establishment probabilities for 23 tree species predicted by the LINKAGES model are presented for the assessment area as a whole, and for each section within the assessment area (Table 39). Early growth potential was also mapped for each species modeled by LINKAGES (Fig. 0). Change in early growth was calculated by dividing the modeled future biomass by the current climate biomass. Change was classified according to the ratios tabulated at right. Modeled:Current biomass <0.4 0.4 through <0.8 0.8 through <1.2 1.2 through <2.0 >2.0 current climate = 0 and future climate model = 0 current climate > 0 and future climate model = 0 Class large decrease small decrease no change small increase large increase not present extirpated Table 39.— Changes in early growth of tree species predicted by the LiNKAGES model for two climate scenarios at the end of the century (2080 through 2099) compared to current climate (1990 through 2009) for 23 species in the assessment area PCM B1 SEPb % change GFDL A1Fi SEPb % change 0.22 0.22 0.26 0.13 0.23 0.24 0.22 0.21 0.26 0.15 0.24 0.23 0.0 -5.5 0.7 14.9 3.5 -5.8 0.02 0.00 0.00 0.00 0.14 0.00 Assessment area 221E 221F M221A M221B M221C 0.14 0.16 0.13 0.07 0.10 0.16 0.16 0.19 0.16 0.12 0.13 0.18 14.3 14.4 23.0 76.3 22.3 12.7 0.18 0.19 0.18 0.14 0.18 0.19 Balsam ir Assessment area 221E 221F M221A M221B M221C 0.00 0.00 0.00 0.00 0.02 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.0 0.0 0.0 0.0 -36.8 0.0 0.00 0.00 0.00 0.00 0.00 0.00 0.0 0.0 0.0 0.0 Exirpated 0.0 Black cherry Assessment area 221E 221F M221A M221B M221C 0.28 0.31 0.30 0.11 0.26 0.32 0.30 0.32 0.32 0.18 0.28 0.32 7.1 1.5 6.4 65.0 8.5 1.8 0.30 0.31 0.32 0.21 0.31 0.31 7.1 -2.1 8.3 93.9 19.9 -1.3 Blackgum Assessment area 221E 221F M221A M221B M221C 0.16 0.18 0.15 0.08 0.12 0.18 0.19 0.21 0.18 0.13 0.14 0.21 18.8 14.7 20.2 75.0 22.6 14.3 0.21 0.21 0.22 0.16 0.21 0.22 31.3 16.8 44.7 109.3 78.1 19.3 Species Seciona American beech Assessment area 221E 221F M221A M221B M221C American elm Current Climate SEPb -90.9 Exirpated Exirpated Exirpated -38.1 Exirpated 28.6 14.4 40.8 111.2 77.7 17.1 (coninued on next page) 282 APPENDiX 4 Table 39 (coninued). PCM B1 SEPb % change GFDL A1Fi SEPb % change 0.19 0.20 0.20 0.11 0.18 0.20 0.20 0.21 0.21 0.16 0.19 0.21 5.3 0.5 2.1 47.6 8.0 0.8 0.13 0.08 0.20 0.15 0.20 0.13 -31.6 -63.1 -0.4 40.2 10.7 -37.5 Assessment area 221E 221F M221A M221B M221C 0.20 0.21 0.21 0.10 0.19 0.21 0.20 0.21 0.21 0.15 0.20 0.21 0.0 0.1 0.0 45.2 5.2 -0.2 0.12 0.06 0.20 0.14 0.20 0.13 -40.0 -72.0 -6.1 32.0 3.3 -37.9 Eastern red cedar Assessment area 221E 221F M221A M221B M221C 0.20 0.23 0.21 0.09 0.15 0.24 0.24 0.26 0.25 0.13 0.19 0.27 20.0 14.7 21.9 43.1 23.6 14.1 0.26 0.26 0.28 0.16 0.26 0.28 30.0 15.3 36.5 67.3 71.4 17.3 Eastern hemlock Assessment area 221E 221F M221A M221B M221C 0.13 0.14 0.15 0.06 0.14 0.15 0.13 0.12 0.15 0.09 0.14 0.13 0.0 -15.2 0.5 43.6 1.4 -13.6 0.01 0.00 0.00 0.00 0.07 0.00 -92.3 Exirpated -99.8 Exirpated -46.7 Exirpated Eastern white pine Assessment area 221E 221F M221A M221B M221C 0.34 0.35 0.38 0.21 0.35 0.37 0.35 0.35 0.38 0.25 0.36 0.38 2.9 1.8 0.0 18.2 2.5 2.1 0.04 0.00 0.01 0.00 0.24 0.00 -88.2 Exirpated -98.3 Exirpated -31.0 Exirpated Flowering dogwood Assessment area 221E 221F M221A M221B M221C 0.05 0.07 0.04 0.03 0.02 0.06 0.08 0.10 0.06 0.05 0.04 0.09 60.0 41.9 66.0 118.1 80.9 45.7 0.10 0.10 0.10 0.07 0.10 0.10 100.0 49.5 162.6 187.9 318.3 61.1 Loblolly pine Assessment area 221E 221F M221A M221B M221C 0.13 0.20 0.03 0.08 0.02 0.17 0.29 0.39 0.17 0.27 0.10 0.35 123.1 94.3 464.8 260.2 455.6 106.1 0.51 0.54 0.50 0.51 0.45 0.53 292.3 167.9 1535.1 577.5 2496.5 212.1 Northern red oak Assessment area 221E 221F M221A M221B M221C 0.29 0.29 0.33 0.17 0.30 0.32 0.31 0.31 0.33 0.20 0.31 0.33 6.9 6.2 0.9 19.5 3.5 3.9 0.18 0.09 0.31 0.18 0.30 0.20 -37.9 -70.4 -5.1 5.5 1.0 -36.4 Species Seciona Black oak Assessment area 221E 221F M221A M221B M221C Chestnut oak Current Climate SEPb (coninued on next page) 283 APPENDiX 4 Table 39 (coninued). PCM B1 SEPb % change GFDL A1Fi SEPb % change 0.35 0.38 0.37 0.15 0.31 0.38 0.37 0.39 0.38 0.25 0.34 0.39 5.7 2.1 5.2 62.0 10.8 1.1 0.36 0.36 0.39 0.28 0.38 0.38 2.9 -6.5 5.8 81.6 23.2 -2.1 Assessment area 221E 221F M221A M221B M221C 0.05 0.07 0.03 0.03 0.02 0.06 0.10 0.13 0.06 0.08 0.04 0.12 100.0 81.7 103.5 182.8 147.2 86.7 0.17 0.18 0.18 0.14 0.15 0.18 240.0 149.8 458.5 401.6 795.6 184.1 Red maple Assessment area 221E 221F M221A M221B M221C 0.31 0.34 0.31 0.24 0.27 0.33 0.34 0.36 0.33 0.32 0.30 0.36 9.7 7.8 9.2 33.4 10.8 7.9 0.37 0.37 0.37 0.35 0.37 0.37 19.4 10.7 22.5 45.3 34.1 12.7 Red spruce Assessment area 221E 221F M221A M221B M221C 0.00 0.00 0.00 0.00 0.02 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.0 0.0 0.0 0.0 -49.9 0.0 0.00 0.00 0.00 0.00 0.00 0.00 0.0 0.0 0.0 0.0 Exirpated 0.0 Scarlet oak Assessment area 221E 221F M221A M221B M221C 0.17 0.18 0.18 0.09 0.17 0.18 0.17 0.18 0.18 0.13 0.17 0.18 0.0 0.5 1.1 44.4 3.2 -0.2 0.04 0.00 0.07 0.02 0.14 0.01 -76.5 -99.7 -60.7 -74.8 -19.6 -96.5 Shortleaf pine Assessment area 221E 221F M221A M221B M221C 0.08 0.14 0.02 0.03 0.01 0.11 0.22 0.31 0.12 0.14 0.07 0.29 175.0 125.3 536.6 305.5 830.0 153.4 0.35 0.36 0.38 0.22 0.34 0.38 337.5 159.2 2017.0 529.6 4515.9 232.5 Sugar maple Assessment area 221E 221F M221A M221B M221C 0.51 0.52 0.52 0.44 0.51 0.52 0.50 0.49 0.52 0.49 0.52 0.51 -2.0 -4.5 -0.1 11.0 0.8 -2.0 0.05 0.00 0.00 0.00 0.33 0.00 -90.2 Exirpated -99.2 -100.0 -34.5 Exirpated Tulip tree Assessment area 221E 221F M221A M221B M221C 0.76 0.77 0.92 0.22 0.78 0.88 0.83 0.85 0.97 0.38 0.84 0.93 9.2 10.9 4.6 69.6 7.1 5.9 0.86 0.85 0.98 0.48 0.91 0.94 Species Seciona Pignut hickory Assessment area 221E 221F M221A M221B M221C Post oak Current Climate SEPb 13.2 11.0 6.1 112.6 15.8 7.0 (coninued on next page) 284 APPENDiX 4 Table 39 (coninued). Current Climate SEPb SEPb Assessment area 221E 221F M221A M221B M221C 0.43 0.37 0.76 0.03 0.50 0.49 0.53 0.49 0.79 0.06 0.56 0.62 23.3 34.2 5.0 88.5 12.8 28.3 0.41 0.25 0.79 0.06 0.60 0.51 -4.7 -32.1 4.6 113.1 20.6 4.5 Assessment area 221E 221F M221A M221B M221C 0.32 0.33 0.36 0.19 0.31 0.36 0.35 0.35 0.38 0.23 0.33 0.37 9.4 7.3 4.9 23.6 8.1 4.5 0.35 0.35 0.38 0.25 0.36 0.37 9.4 5.7 5.5 34.4 17.6 4.9 Species Seciona White ash White oak PCM B1 % change GFDL A1Fi SEPb % change Assessment area values were derived from the weighted average of secions. Species establishment probabiliies (SEP) are derived from LINKAGES model results. SEP is a value relecing the ability of a species to establish and grow on a site, is scaled 0 through 1, and is relaive to the other species considered. It is important to look at absolute and percentage changes together; in some cases a small absolute change can result in a large percentage change. a b A forest road on the Monongahela Naional Forest, West Virginia. Photo by Patricia Butler, NIACS and Michigan Tech, used with permission. 285 APPENDiX 4 Figure 60.—Changes in early growth of tree species under two climate scenarios for the end of the century (2080 through 2099) compared to current climate (1990 through 2009). Change is based on predicted biomass by the LINKAGES model ater 30 years of establishment and growth from bare ground and calculated as predicted biomass for each future climate scenario divided by predicted biomass under current climate, and then put into categories. 28 APPENDiX 4 Figure 60 (coninued). 287 APPENDiX 4 Figure 60 (coninued). 288 APPENDiX 4 Figure 60 (coninued). LANDiS PRo MoDEL RESuLTS In contrast to predictions by LINKAGES, LANDIS PRO simulates stand- and landscape-level processes such as competition, management, seed dispersal, and disturbance. In the scenarios below, however, these factors were held constant among model simulations, so that differences among current climate and future climate scenarios are the result of the effects of precipitation and temperature on species basal area (square feet per acre) and trees per acre. Percentage change over time within a climate scenario (columns) shows how current species are predicted to change under that scenario (Tables 40 through 42). The relative differences in the values for percentage change across scenarios indicate the differences between climate scenarios. It is important to consider both the absolute and the percentage change, especially if considering multiple species. Percentage changes are relative only to a particular species and may exaggerate a projected change. Figures 1 through 4 present these changes in basal area and trees per acre for PCM B1 and GFDL A1FI. The width of a line represents the species’ relative abundance; over time the width of the line increases or decreases in response to climate variables. REFERENCES Matthews, S.N.; Iverson, L.R.; Prasad, A.M.; Peters, M.P.; Rodewald, P.G. 2011. Modifying climate change habitat models using tree speciesspecific assessments of model uncertainty and life history-factors. Forest Ecology and Management. 22(8): 140-1472. 289 Current climate Change BA in BA in from 2009 2040 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2040 climatec 2040 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2040 2009b 2040 climatec 2040 climatec Tree species Seciona American beech Assessment area 221E 221F M221A M221B M221C 5.1 3.9 4.5 7.2 6.6 6.1 4.4 3.8 4.0 5.4 4.9 5.0 -15% -3% -11% -25% -26% -18% 4.4 3.8 4.1 5.7 5.0 5.0 1% 0% 2% 6% 2% 0% 4.3 3.6 3.9 5.6 5.0 4.8 -2% -5% -3% 4% 2% -4% 26.8 19.7 16.7 39.8 36.0 36.2 15.5 13.3 15.3 18.1 17.9 18.2 -42% -32% -8% -55% -50% -50% 15.7 13.6 15.2 18.8 17.9 18.1 1% 2% -1% 4% 0% -1% 12.8 10.5 12.8 15.0 15.5 15.2 -18% -21% -16% -17% -13% -16% Black cherry Assessment area 221E 221F M221A M221B M221C 7.3 7.2 13.8 8.0 7.4 3.6 9.0 9.1 16.7 8.8 9.1 4.3 23% 26% 21% 10% 23% 19% 9.0 9.1 16.8 8.9 9.1 4.3 0% 0% 1% 1% 0% 0% 9.4 9.4 17.6 9.1 9.4 4.6 4% 3% 5% 3% 3% 7% 18.9 19.3 42.3 15.8 15.4 10.8 12.8 13.5 34.2 8.4 8.9 5.6 -33% -30% -19% -47% -42% -48% 12.9 13.7 34.4 8.7 9.0 5.6 1% 1% 1% 4% 1% 0% 11.5 12.1 30.9 7.6 8.2 4.9 -10% -10% -10% -10% -8% -13% Black oak Assessment area 221E 221F M221A M221B M221C 6.8 7.3 2.1 6.4 7.0 8.0 4.8 5.2 1.4 4.2 5.1 5.4 -30% -29% -33% -34% -27% -33% 4.8 5.2 1.4 4.2 5.1 5.4 0% 0% 0% 0% 0% 0% 3.4 3.6 0.8 3.0 3.7 3.8 -30% -31% -43% -29% -27% -30% 8.4 9.5 1.5 8.9 9.0 7.7 3.9 4.4 1.1 3.7 4.0 3.9 -54% -54% -27% -58% -56% -49% 4.0 4.5 1.1 3.9 4.0 4.0 2% 2% 0% 5% 0% 3% 3.1 3.5 0.8 3.1 3.2 3.0 -21% -20% -27% -16% -20% -23% Chestnut oak Assessment area 221E 221F M221A M221B M221C 5.6 4.2 0.2 8.3 8.3 8.2 4.5 3.6 0.1 6.5 6.5 6.1 -20% -14% -50% -22% -22% -26% 4.6 3.6 0.1 6.9 6.6 6.2 2% 0% 0% 6% 2% 2% 4.4 3.4 0.1 6.8 6.7 5.8 -2% -6% 0% 5% 3% -5% 14.9 9.6 0.9 21.7 24.2 24.1 25.8 20.1 1.8 31.6 41.0 35.6 73% 109% 100% 46% 69% 48% 27.0 21.2 1.6 36.2 41.2 36.8 5% 5% -11% 15% 0% 3% 24.4 16.5 1.7 36.2 42.9 32.8 -5% -18% -6% 15% 5% -8% Eastern hemlock Assessment area 221E 221F M221A M221B M221C 1.7 0.5 0.2 3.2 3.4 3.1 1.2 0.4 0.2 2.1 2.3 2.3 -29% -20% 0% -34% -32% -26% 1.2 0.4 0.2 2.2 2.3 2.3 1% 0% 0% 5% 0% 0% 1.1 0.4 0.2 2.0 2.2 2.1 -5% 0% 0% -5% -4% -9% 6.1 1.7 0.7 9.4 11.0 15.3 3.2 1.0 1.1 5.2 6.2 6.7 -47% -41% 57% -45% -44% -56% 3.3 1.1 1.0 5.5 6.3 6.8 3% 10% -9% 6% 2% 1% 2.5 0.8 0.7 3.9 5.1 5.2 -21% -20% -36% -25% -18% -22% Eastern redcedar Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.2 0.1 0.0 0.0 -56% 0.0 -100% 0.0 0% 0.2 0% 0.1 0% 0.0 0% 0.0 0.0 0.0 0.2 0.1 0.0 0% 0% 0% 0% 0% 0% 0.0 0.0 0.0 0.2 0.1 0.0 0% 0% 0% 0% 0% 0% 0.8 0.7 0.0 1.6 1.2 0.4 0.8 0.4 0.0 2.9 1.4 0.4 2% -43% 0% 81% 17% 0% 0.8 0.4 0.0 3.2 1.3 0.4 0% 0% 0% 10% -7% 0% 0.5 0.2 0.0 1.8 0.9 0.2 -41% -50% 0% -38% -36% -50% (coninued on next page) APPENDiX 4 290 Table 40.—Absolute and percentage change in basal area (BA) and trees per acre (TPA) predicted by the LANDIS PRO model for 17 species for current climate and two climate scenarios for the assessment area in year 2040 Table 40 (coninued). Current climate BA in 2009 Change BA in from 2040 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2040 climatec 2040 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2040 2009b 2040 climatec 2040 climatec Seciona Eastern whitepine Assessment area 221E 221F M221A M221B M221C 1.2 0.9 0.4 1.4 1.9 1.4 1.2 1.0 0.4 1.8 2.2 1.0 7% 11% 0% 29% 16% -29% 1.2 1.0 0.4 1.8 2.2 1.0 0% 0% 0% 0% 0% 0% 1.2 0.9 0.4 1.7 2.1 0.9 -7% -10% 0% -6% -5% -10% 7.3 4.7 1.3 12.8 14.7 6.8 6.6 4.8 4.1 11.1 12.8 3.2 -10% 2% 215% -13% -13% -53% 6.8 5.0 4.1 12.0 12.7 3.4 3% 4% 0% 8% -1% 6% 4.6 3.2 2.5 7.8 9.6 2.1 -30% -33% -39% -30% -25% -34% Loblolly pine Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.1 0.1 0.0 0.1 39% 0.2 100% 0.0 0% 0.1 0% 0.0 -100% 0.0 0% 0.1 0.2 0.0 0.1 0.0 0.0 0% 0% 0% 0% 0% 0% 0.1 0.2 0.0 0.1 0.1 0.0 17% 0% 0% 0% NA 0% 0.5 0.8 0.0 0.4 0.3 0.0 0.3 0.6 0.0 0.1 0.1 0.0 -34% -25% 0% -75% -67% 0% 0.4 0.7 0.0 0.1 0.1 0.0 15% 17% 0% 0% 0% 0% 0.3 0.6 0.0 0.1 0.1 0.0 0% 0% 0% 0% 0% 0% Northern red oak Assessment area 221E 221F M221A M221B M221C 6.8 5.2 5.9 8.7 9.3 7.9 7.7 6.1 6.1 9.4 10.8 9.1 14% 17% 3% 8% 16% 15% 7.8 6.1 6.3 9.5 10.9 9.2 1% 0% 3% 1% 1% 1% 8.0 6.3 6.2 9.8 11.2 9.2 3% 3% 2% 4% 4% 1% 11.9 10.2 10.9 13.6 16.6 11.3 6.8 6.1 6.2 7.6 8.2 6.9 -43% -40% -43% -44% -51% -39% 7.5 6.5 8.3 8.5 9.2 7.4 11% 7% 34% 12% 12% 7% 6.9 5.9 8.2 7.7 8.6 6.7 2% -3% 32% 1% 5% -3% Pignut hickory Assessment area 221E 221F M221A M221B M221C 2.6 2.6 0.5 2.9 2.9 3.0 2.5 2.6 0.4 2.9 2.9 2.8 -2% 0% -20% 0% 0% -7% 2.6 2.6 0.4 3.0 3.0 2.8 1% 0% 0% 3% 3% 0% 2.6 2.6 0.4 3.0 3.1 2.7 1% 0% 0% 3% 7% -4% 7.3 6.9 0.7 10.3 9.8 7.6 6.3 6.4 1.0 7.2 8.3 6.1 -14% -7% 43% -30% -15% -20% 6.5 6.6 1.0 7.9 8.3 6.3 3% 3% 0% 10% 0% 3% 5.9 5.7 1.0 7.1 8.0 5.7 -7% -11% 0% -1% -4% -7% Red maple Assessment area 221E 221F M221A M221B M221C 16.0 13.8 23.6 18.0 17.5 15.5 22.7 21.0 32.3 22.4 23.2 22.3 42% 52% 37% 24% 33% 44% 22.9 21.1 32.6 22.8 23.5 22.4 1% 0% 1% 2% 1% 0% 24.6 22.9 34.2 24.0 25.0 24.4 8% 9% 6% 7% 8% 9% 77.4 72.7 81.6 66.8 77.0 97.2 61.4 60.5 110.8 50.9 49.0 57.7 -21% -17% 36% -24% -36% -41% 62.8 62.0 111.7 55.5 49.4 58.6 2% 2% 1% 9% 1% 2% 59.0 57.2 109.1 53.8 47.9 52.8 -4% -5% -2% 6% -2% -8% Red spruce Assessment area 221E 221F M221A M221B M221C 1.0 0.3 0.0 3.4 2.5 0.3 0.5 0.1 0.0 1.7 1.4 0.1 -50% -67% 0% -50% -44% -67% 0.5 0.1 0.0 1.8 1.4 0.1 2% 0% 0% 6% 0% 0% 0.5 0.1 0.0 1.6 1.3 0.1 -6% 0% 0% -6% -7% 0% 5.1 1.7 0.0 17.0 13.0 1.4 1.3 0.2 0.0 3.2 4.8 0.2 -74% -88% 0% -81% -63% -86% 1.3 0.2 0.0 3.3 4.6 0.2 0% 0% 0% 3% -4% 0% 1.0 0.1 0.0 2.5 3.6 0.1 0% -50% 0% -22% -25% -50% (coninued on next page) APPENDiX 4 291 Tree species Current climate BA in 2009 Change BA in from 2040 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2040 climatec 2040 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2040 2009b 2040 climatec 2040 climatec Tree species Seciona Scarlet oak Assessment area 221E 221F M221A M221B M221C 4.8 3.4 0.3 6.7 7.4 7.8 3.5 2.0 0.3 5.5 6.3 5.6 -27% -41% 0% -18% -15% -28% 3.5 2.0 0.3 5.4 6.3 5.6 0% 0% 0% -2% 0% 0% 2.4 1.2 0.3 4.1 4.5 3.5 -33% -40% 0% -25% -29% -38% 4.0 2.5 0.5 6.4 6.7 6.2 2.3 1.4 0.4 3.5 3.9 3.5 -43% -44% -20% -45% -42% -44% 2.3 1.4 0.4 3.6 4.0 3.6 2% 0% 0% 3% 3% 3% 1.8 1.0 0.4 2.8 3.2 2.6 -23% -29% 0% -20% -18% -26% Sugar maple Assessment area 221E 221F M221A M221B M221C 6.1 5.9 7.0 6.8 6.4 5.5 7.2 7.3 8.4 7.4 6.9 6.1 17% 24% 20% 9% 8% 11% 7.3 7.4 8.5 7.9 7.0 6.2 2% 1% 1% 7% 1% 2% 7.1 7.1 8.3 7.7 7.1 6.0 -1% -3% -1% 4% 3% -2% 54.1 53.4 73.8 47.2 49.9 54.7 54.4 55.9 73.4 62.1 48.0 41.4 0% 5% -1% 32% -4% -24% 55.7 57.7 72.3 67.8 47.7 41.9 2% 3% -1% 9% -1% 1% 39.6 38.1 51.3 49.4 39.3 31.8 -27% -32% -30% -20% -18% -23% Tulip tree Assessment area 221E 221F M221A M221B M221C 6.5 6.6 2.1 5.8 5.7 9.9 8.4 9.0 3.5 6.1 7.2 12.5 31% 36% 67% 5% 26% 26% 8.6 9.1 3.5 6.3 7.3 12.7 1% 1% 0% 3% 1% 2% 8.7 9.1 3.8 6.3 7.7 13.0 3% 1% 9% 3% 7% 4% 22.9 25.0 9.5 14.1 16.9 36.6 67.3 76.0 62.1 21.3 53.3 88.7 194% 204% 554% 51% 215% 142% 69.6 79.5 61.1 24.1 52.4 92.4 3% 5% -2% 13% -2% 4% 68.5 75.7 59.0 22.4 57.8 93.1 2% 0% -5% 5% 8% 5% White ash Assessment area 221E 221F M221A M221B M221C 2.5 2.7 4.8 1.9 1.7 1.7 2.8 3.1 5.5 2.1 2.0 1.8 14% 15% 15% 11% 18% 6% 2.8 3.1 5.6 2.1 2.0 1.8 0% 0% 2% 0% 0% 0% 2.8 3.0 5.3 2.1 2.1 1.8 -2% -3% -4% 0% 5% 0% 11.9 14.0 19.5 8.1 8.0 8.0 10.9 10.4 35.9 3.8 8.5 5.9 -8% -26% 84% -53% 6% -26% 11.3 11.0 36.3 4.1 8.6 6.1 4% 6% 1% 8% 1% 3% 9.6 8.6 32.9 3.4 8.1 5.1 -12% -17% -8% -11% -5% -14% White oak Assessment area 221E 221F M221A M221B M221C 5.7 6.6 1.1 5.6 5.6 5.7 5.9 7.1 1.4 5.2 5.4 5.6 3% 8% 27% -7% -4% -2% 6.0 7.2 1.4 5.5 5.5 5.7 2% 1% 0% 6% 2% 2% 6.0 7.0 1.4 5.7 5.8 5.8 2% -1% 0% 10% 7% 4% 12.8 15.0 2.9 12.8 12.3 12.2 36.6 43.2 13.8 29.8 34.5 35.3 185% 188% 376% 133% 180% 189% 38.1 45.2 14.1 33.0 34.9 36.2 4% 5% 2% 11% 1% 3% 35.2 40.0 12.8 30.2 35.0 35.9 -4% -7% -7% 1% 1% 2% Assessment area values were derived from the weighted average of secions. Change under current climate represents the diference from 2009 through 2100 due to succession and management, but not climate. c Change from current climate for PCM B1 and GFDL A1FI represents the diference between these scenarios and current climate in 2100 and represents the potenial change due to climate change. a b APPENDiX 4 292 Table 40 (coninued). Table 41.—Absolute and percentage change in basal area (BA) and trees per acre (TPA) predicted by the LANDIS PRO model for 17 species under current climate and two climate scenarios for the assessment area in year 2070 Tree species Seciona AmerAmerican beech Assessment area 221E 221F M221A M221B M221C Current climate Change BA in BA in from 2009 2070 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2070 climatec 2070 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2070 2009b 2070 climatec 2070 climatec 5.1 3.9 4.5 7.2 6.6 6.1 4.4 3.8 3.8 5.6 5.1 5.1 -14% -3% -16% -22% -23% -16% 4.4 3.8 3.9 5.9 5.1 5.1 1% 0% 3% 5% 0% 0% 4.3 3.6 3.7 5.6 5.1 4.9 -3% -5% -3% 0% 0% -4% 26.8 19.7 16.7 39.8 36.0 36.2 12.9 11.0 11.8 14.7 15.8 14.9 -52% -44% -29% -63% -56% -59% 12.7 10.7 11.9 15.7 15.7 14.2 -1% -3% 1% 7% -1% -5% 10.3 8.1 9.1 12.1 14.5 12.1 -20% -26% -23% -18% -8% -19% Assessment area 221E 221F M221A M221B M221C 7.3 7.2 13.8 8.0 7.4 3.6 7.6 7.8 15.6 6.4 7.1 4.1 4% 8% 13% -20% -4% 14% 7.8 8.0 15.5 6.4 7.1 4.2 1% 3% -1% 0% 0% 2% 7.7 7.9 15.6 6.5 7.0 4.1 1% 1% 0% 2% -1% 0% 18.9 19.3 42.3 15.8 15.4 10.8 9.4 10.0 26.5 5.3 6.2 4.0 -51% -48% -37% -66% -60% -63% 9.5 10.0 26.6 5.6 6.4 4.1 1% 0% 0% 6% 3% 2% 9.2 9.7 26.0 5.6 6.4 4.0 -1% -3% -2% 6% 3% 0% Black oak Assessment area 221E 221F M221A M221B M221C 6.8 7.3 2.1 6.4 7.0 8.0 2.1 2.4 0.4 1.9 2.3 2.1 -69% -67% -81% -70% -67% -74% 2.2 2.5 0.4 2.0 2.3 2.1 3% 4% 0% 5% 0% 0% 2.1 2.4 0.4 2.0 2.4 2.1 1% 0% 0% 5% 4% 0% 8.4 9.5 1.5 8.9 9.0 7.7 2.4 2.8 0.6 2.2 2.3 2.2 -72% -71% -60% -75% -74% -71% 2.4 2.8 0.6 2.4 2.4 2.2 2% 0% 0% 9% 4% 0% 2.3 2.6 0.6 2.3 2.4 2.1 -4% -7% 0% 5% 4% -5% Chestnut oak Assessment area 221E 221F M221A M221B M221C 5.6 4.2 0.2 8.3 8.3 8.2 4.7 3.6 0.2 6.8 7.1 6.4 -17% -14% 0% -18% -14% -22% 4.8 3.7 0.2 7.5 7.2 6.5 3% 3% 0% 10% 1% 2% 4.5 3.3 0.2 7.3 7.2 6.0 -3% -8% 0% 7% 1% -6% 14.9 9.6 0.9 21.7 24.2 24.1 29.4 22.4 1.4 36.5 48.4 40.0 97% 133% 56% 68% 100% 66% 31.2 22.6 1.4 46.8 52.0 40.8 6% 1% 0% 28% 7% 2% 25.5 14.6 1.3 44.2 50.2 32.4 -13% -35% -7% 21% 4% -19% Eastern hemlock Assessment area 221E 221F M221A M221B M221C 1.7 0.5 0.2 3.2 3.4 3.1 1.1 0.4 0.2 1.9 2.1 2.1 -34% -20% 0% -41% -38% -32% 1.1 0.4 0.2 2.0 2.1 2.1 1% 0% 0% 5% 0% 0% 1.1 0.4 0.2 1.9 2.1 2.0 -1% 0% 0% 0% 0% -5% 6.1 1.7 0.7 9.4 11.0 15.3 2.9 1.1 1.2 4.0 5.5 5.8 -52% -35% 71% -57% -50% -62% 2.8 1.1 1.3 4.4 5.4 5.1 -3% 0% 8% 10% -2% -12% 2.1 0.7 0.6 2.9 4.7 3.9 -27% -36% -50% -28% -15% -33% Eastern redcedar Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.2 0.1 0.0 0.0 -56% 0.0 -100% 0.0 0% 0.2 0% 0.1 0% 0.0 0% 0.0 0.0 0.0 0.2 0.1 0.0 0% 0% 0% 0% 0% 0% 0.0 0.0 0.0 0.2 0.1 0.0 0% 0% 0% 0% 0% 0% 0.8 0.7 0.0 1.6 1.2 0.4 0.2 0.1 0.0 0.8 0.3 0.1 -75% -86% 0% -50% -75% 0% 0.2 0.1 0.0 0.9 0.3 0.1 0% 0% 0% 13% 0% 0% 0.2 0.1 0.0 0.8 0.3 0.1 0% 0% 0% 0% 0% 0% (coninued on next page) APPENDiX 4 293 Black cherry Current climate BA in 2009 Change BA in from 2070 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2070 climatec 2070 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2070 2009b 2070 climatec 2070 climatec Tree species Seciona Eastern white pine Assessment area 221E 221F M221A M221B M221C 1.2 0.9 0.4 1.4 1.9 1.4 1.1 0.9 0.4 1.7 2.0 0.9 -2% 0% 0% 21% 5% -36% 1.1 0.9 0.4 1.8 2.0 0.9 1% 0% 0% 6% 0% 0% 1.1 0.8 0.4 1.6 2.0 0.8 -6% -11% 0% -6% 0% -11% 7.3 4.7 1.3 12.8 14.7 6.8 3.7 2.9 2.0 6.7 6.8 1.7 -49% -38% 54% -48% -54% -75% 3.7 2.8 2.1 7.2 6.9 1.7 1% -3% 5% 7% 1% 0% 2.7 1.8 1.2 4.8 6.2 1.1 -26% -38% -40% -28% -9% -35% Loblolly pine Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.1 0.1 0.0 0.2 0.3 0.0 0.1 0.1 0.0 127% 200% 0% 0% 0% 0% 0.2 0.3 0.0 0.1 0.1 0.0 0% 0% 0% 0% 0% 0% 0.2 0.3 0.0 0.1 0.1 0.0 0% 0% 0% 0% 0% 0% 0.5 0.8 0.0 0.4 0.3 0.0 0.2 0.4 0.0 0.1 0.1 0.0 -54% -50% 0% -75% -67% 0% 0.3 0.5 0.0 0.1 0.1 0.0 22% 25% 0% 0% 0% 0% 0.3 0.5 0.0 0.1 0.1 0.0 22% 25% 0% 0% 0% 0% Northern red oak Assessment area 221E 221F M221A M221B M221C 6.8 5.2 5.9 8.7 9.3 7.9 5.8 4.7 4.5 6.6 7.8 6.9 -15% -10% -24% -24% -16% -13% 5.9 4.9 4.8 6.7 8.0 6.9 3% 4% 7% 2% 3% 0% 5.9 4.8 4.9 6.7 8.1 6.8 2% 2% 9% 2% 4% -1% 11.9 10.2 10.9 13.6 16.6 11.3 4.9 4.9 4.0 5.2 4.9 4.9 -59% -52% -63% -62% -70% -57% 6.4 5.7 8.4 6.8 7.6 5.9 32% 16% 110% 31% 55% 20% 5.9 4.9 8.1 6.5 7.6 5.4 22% 0% 103% 25% 55% 10% Pignut hickory Assessment area 221E 221F M221A M221B M221C 2.6 2.6 0.5 2.9 2.9 3.0 2.7 2.7 0.4 3.0 3.2 3.0 4% 4% -20% 3% 10% 0% 2.7 2.8 0.4 3.1 3.2 3.0 2% 4% 0% 3% 0% 0% 2.7 2.7 0.4 3.1 3.2 2.9 0% 0% 0% 3% 0% -3% 7.3 6.9 0.7 10.3 9.8 7.6 5.6 5.8 1.0 5.5 7.2 5.6 -24% -16% 43% -47% -27% -26% 5.8 5.8 1.0 6.6 7.7 5.9 4% 0% 0% 20% 7% 5% 5.6 5.4 1.0 6.6 7.7 5.4 0% -7% 0% 20% 7% -4% Red maple Assessment area 221E 221F M221A M221B M221C 16.0 13.8 23.6 18.0 17.5 15.5 27.6 26.6 35.9 24.8 26.7 28.8 73% 93% 52% 38% 53% 86% 27.9 27.0 35.9 25.7 26.6 29.1 1% 2% 0% 4% 0% 1% 27.6 26.4 35.8 25.8 27.1 28.4 0% -1% 0% 4% 1% -1% 77.4 72.7 81.6 66.8 77.0 97.2 58.2 58.7 108.7 51.7 44.8 48.6 -25% -19% 33% -23% -42% -50% 60.5 59.3 114.5 61.1 46.1 50.9 4% 1% 5% 18% 3% 5% 59.6 57.6 113.0 59.8 48.2 49.4 2% -2% 4% 16% 8% 2% Red spruce Assessment area 221E 221F M221A M221B M221C 1.0 0.3 0.0 3.4 2.5 0.3 0.4 0.1 0.0 1.4 1.2 0.1 -57% -67% 0% -59% -52% -67% 0.4 0.1 0.0 1.4 1.2 0.1 0% 0% 0% 0% 0% 0% 0.4 0.1 0.0 1.4 1.2 0.1 0% 0% 0% 0% 0% 0% 5.1 1.7 0.0 17.0 13.0 1.4 0.7 0.1 0.0 1.5 2.8 0.1 -86% -94% 0% -91% -78% -93% 0.7 0.1 0.0 1.5 2.7 0.1 0% 0% 0% 0% -4% 0% 0.6 0.1 0.0 1.4 2.2 0.1 0% 0% 0% -7% -21% 0% (coninued on next page) APPENDiX 4 294 Table 41 (coninued). Table 41 (coninued). Current climate BA in 2009 Change BA in from 2070 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2070 climatec 2070 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2070 2009b 2070 climatec 2070 climatec Tree species Seciona Scarlet oak Assessment area 221E 221F M221A M221B M221C 4.8 3.4 0.3 6.7 7.4 7.8 0.9 0.5 0.2 1.5 1.6 1.5 -81% -85% -33% -78% -78% -81% 0.9 0.5 0.2 1.5 1.6 1.5 0% 0% 0% 0% 0% 0% 0.9 0.5 0.2 1.5 1.6 1.4 -2% 0% 0% 0% 0% -7% 4.0 2.5 0.5 6.4 6.7 6.2 1.3 0.8 0.4 1.8 2.2 2.0 -68% -68% -20% -72% -67% -68% 1.3 0.8 0.4 2.0 2.2 2.1 3% 0% 0% 11% 0% 5% 1.2 0.7 0.3 1.7 2.1 1.7 -10% -13% -25% -6% -5% -15% Sugar maple Assessment area 221E 221F M221A M221B M221C 6.1 5.9 7.0 6.8 6.4 5.5 7.5 7.6 8.5 8.1 7.3 6.4 23% 29% 21% 19% 14% 16% 7.6 7.7 8.6 8.5 7.3 6.4 1% 1% 1% 5% 0% 0% 7.2 7.2 8.3 8.0 7.3 6.1 -4% -5% -2% -1% 0% -5% 54.1 53.4 73.8 47.2 49.9 54.7 43.2 42.0 44.7 63.3 42.9 33.6 -20% -21% -39% 34% -14% -39% 42.0 40.5 44.4 64.3 41.3 31.9 -3% -4% -1% 2% -4% -5% 30.3 27.5 30.2 41.6 37.6 23.1 -30% -35% -32% -34% -12% -31% Tulip tree Assessment area 221E 221F M221A M221B M221C 6.5 6.6 2.1 5.8 5.7 9.9 9.9 10.4 4.9 6.0 8.7 14.9 53% 58% 133% 3% 53% 51% 10.0 10.6 4.9 6.3 8.8 15.0 2% 2% 0% 5% 1% 1% 9.8 10.2 4.9 6.5 8.9 14.6 -1% -2% 0% 8% 2% -2% 22.9 25.0 9.5 14.1 16.9 36.6 80.2 91.0 83.1 19.9 65.4 100.6 251% 264% 775% 41% 287% 175% 84.6 94.4 84.6 24.8 70.5 108.7 5% 4% 2% 25% 8% 8% 81.2 88.6 83.1 27.2 72.8 101.1 1% -3% 0% 37% 11% 0% White ash Assessment area 221E 221F M221A M221B M221C 2.5 2.7 4.8 1.9 1.7 1.7 3.0 3.3 5.3 2.2 2.2 2.0 20% 22% 10% 16% 29% 18% 3.0 3.3 5.2 2.3 2.3 2.1 1% 0% -2% 5% 5% 5% 2.9 3.2 5.2 2.2 2.3 2.0 -1% -3% -2% 0% 5% 0% 11.9 14.0 19.5 8.1 8.0 8.0 8.7 8.0 30.9 2.3 7.2 4.4 -27% -43% 58% -72% -10% -45% 9.2 8.7 30.5 2.6 7.8 4.8 6% 9% -1% 13% 8% 9% 8.2 6.9 29.8 2.6 7.8 4.3 -6% -14% -4% 13% 8% -2% White oak Assessment area 221E 221F M221A M221B M221C 5.7 6.6 1.1 5.6 5.6 5.7 6.9 8.1 1.5 6.3 6.7 6.9 21% 23% 36% 13% 20% 21% 7.1 8.3 1.6 6.7 6.8 7.0 3% 2% 7% 6% 1% 1% 6.8 7.8 1.5 6.5 6.8 6.7 -2% -4% 0% 3% 1% -3% 12.8 15.0 2.9 12.8 12.3 12.2 48.5 56.8 18.4 38.2 47.5 46.8 278% 279% 534% 198% 286% 284% 51.1 58.4 18.9 46.1 51.2 49.2 5% 3% 3% 21% 8% 5% 48.1 53.8 17.8 43.9 51.0 46.2 -1% -5% -3% 15% 7% -1% Assessment area values were derived from the weighted average of secions. Change under current climate represents the diference from 2009 through 2100 due to succession and management, but not climate. c Change from current climate for PCM B1 and GFDL A1FI represents the diference between these scenarios and current climate in 2100 and represents the potenial change due to climate change. a 295 APPENDiX 4 b Current climate Change BA in BA in from 2009 2100 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2100 climatec 2100 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2100 2009b 2100 climatec 2100 climatec Tree species Seciona American beech Assessment area 221E 221F M221A M221B M221C 5.1 3.9 4.5 7.2 6.6 6.1 4.6 4.0 3.6 5.7 5.5 5.4 -9% 3% -20% -21% -17% -11% 4.6 4.0 3.7 6.0 5.5 5.3 0% 0% 3% 5% 0% -2% 4.2 3.5 3.2 5.3 5.3 4.8 -10% -13% -11% -7% -4% -11% 26.8 19.7 16.7 39.8 36.0 36.2 14.2 11.4 12.7 14.7 17.9 19.0 -47% -42% -24% -63% -50% -48% 13.9 11.0 13.2 15.3 18.1 17.4 -2% -4% 4% 4% 1% -8% 8.1 5.9 5.4 8.2 14.0 9.2 -43% -48% -57% -44% -22% -52% Black cherry Assessment area 221E 221F M221A M221B M221C 7.3 7.2 13.8 8.0 7.4 3.6 4.6 4.8 10.6 3.0 3.5 2.7 -38% -33% -23% -63% -53% -25% 4.6 4.9 10.6 3.0 3.5 2.8 1% 2% 0% 0% 0% 4% 4.5 4.7 10.5 3.0 3.5 2.7 -1% -2% -1% 0% 0% 0% 18.9 19.3 42.3 15.8 15.4 10.8 7.2 7.6 21.8 3.2 4.6 3.2 -62% -61% -48% -80% -70% -70% 7.3 7.7 21.7 3.8 4.9 3.2 2% 1% 0% 19% 7% 0% 7.3 7.6 21.3 4.0 5.1 3.2 2% 0% -2% 25% 11% 0% Black oak Assessment area 221E 221F M221A M221B M221C 6.8 7.3 2.1 6.4 7.0 8.0 1.5 1.7 0.3 1.5 1.6 1.6 -78% -77% -86% -77% -77% -80% 1.5 1.7 0.3 1.6 1.6 1.6 1% 0% 0% 7% 0% 0% 1.5 1.6 0.3 1.6 1.7 1.5 -2% -6% 0% 7% 6% -6% 8.4 9.5 1.5 8.9 9.0 7.7 2.0 2.3 0.5 1.7 2.0 1.8 -77% -76% -67% -81% -78% -77% 2.0 2.3 0.5 1.9 2.0 1.8 1% 0% 0% 12% 0% 0% 1.8 1.9 0.5 1.9 2.1 1.6 -10% -17% 0% 12% 5% -11% Chestnut oak Assessment area 221E 221F M221A M221B M221C 5.6 4.2 0.2 8.3 8.3 8.2 4.9 3.6 0.2 7.4 7.7 6.8 -12% -14% 0% -11% -7% -17% 5.1 3.7 0.2 8.2 7.8 6.8 3% 3% 0% 11% 1% 0% 4.6 3.1 0.2 8.1 7.8 6.1 -5% -14% 0% 9% 1% -10% 14.9 9.6 0.9 21.7 24.2 24.1 30.9 21.7 1.5 43.6 53.5 41.6 108% 126% 67% 101% 121% 73% 32.6 21.8 1.6 58.3 55.1 41.1 5% 0% 7% 34% 3% -1% 25.0 10.0 1.5 60.9 53.7 28.6 -19% -54% 0% 40% 0% -31% Eastern hemlock Assessment area 221E 221F M221A M221B M221C 1.7 0.5 0.2 3.2 3.4 3.1 1.1 0.4 0.2 1.8 2.1 2.0 -35% -20% 0% -44% -38% -35% 1.1 0.4 0.2 1.9 2.1 2.0 1% 0% 0% 6% 0% 0% 1.0 0.3 0.2 1.8 2.0 1.8 -9% -25% 0% 0% -5% -10% 6.1 1.7 0.7 9.4 11.0 15.3 3.8 1.4 1.5 4.2 8.1 7.3 -38% -18% 114% -55% -26% -52% 3.6 1.3 1.5 4.8 7.8 6.1 -6% -7% 0% 14% -4% -16% 2.0 0.6 0.5 2.2 5.7 2.9 -47% -57% -67% -48% -30% -60% Eastern redcedar Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.2 0.1 0.0 0.0 -89% 0.0 -100% 0.0 0% 0.1 -50% 0.0 -100% 0.0 0% 0.0 0.0 0.0 0.1 0.0 0.0 0% 0% 0% 0% 0% 0% 0.0 0.0 0.0 0.1 0.0 0.0 0% 0% 0% 0% 0% 0% 0.8 0.7 0.0 1.6 1.2 0.4 0.1 -91% 0.0 -100% 0.0 0% 0.5 -69% 0.1 -92% 0.0 -100% 0.0 0.0 0.0 0.3 0.1 0.0 0% 0% 0% -40% 0% 0% 0.1 0.1 0.0 0.6 0.1 0.0 87% NA 0% 20% 0% 0% (coninued on next page) APPENDiX 4 29 Table 42.—Absolute and percent change in basal area (BA) and trees per acre (TPA) predicted by the LANDIS PRO model for 17 species for current climate and two climate scenarios for the assessment area in year 2100 Table 42 (coninued). Current climate BA in 2009 Change BA in from 2100 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2100 climatec 2100 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2100 2009b 2100 climatec 2100 climatec Seciona Eastern white pine Assessment area 221E 221F M221A M221B M221C 1.2 0.9 0.4 1.4 1.9 1.4 1.0 0.8 0.3 1.6 1.8 0.8 -13% -11% -25% 14% -5% -43% 1.0 0.8 0.3 1.6 1.8 0.8 0% 0% 0% 0% 0% 0% 0.9 0.7 0.3 1.4 1.8 0.8 -7% -13% 0% -13% 0% 0% 7.3 4.7 1.3 12.8 14.7 6.8 Loblolly pine Assessment area 221E 221F M221A M221B M221C 0.1 0.1 0.0 0.1 0.1 0.0 0.2 0.3 0.0 0.1 0.1 0.0 127% 200% 0% 0% 0% 0% 0.2 0.3 0.0 0.1 0.1 0.0 0% 0% 0% 0% 0% 0% 0.2 0.3 0.0 0.1 0.1 0.0 0% 0% 0% 0% 0% 0% 0.5 0.8 0.0 0.4 0.3 0.0 Northern red oak Assessment area 221E 221F M221A M221B M221C 6.8 5.2 5.9 8.7 9.3 7.9 2.8 2.8 2.2 2.8 3.2 2.8 -58% -46% -63% -68% -66% -65% 3.3 3.1 3.1 3.3 3.9 3.0 15% 11% 41% 18% 22% 7% 3.1 2.8 3.1 3.2 3.9 2.8 9% 0% 41% 14% 22% 0% Pignut hickory Assessment area 221E 221F M221A M221B M221C 2.6 2.6 0.5 2.9 2.9 3.0 2.7 2.7 0.3 2.8 3.3 3.2 5% 4% -40% -3% 14% 7% 2.8 2.8 0.3 3.0 3.4 3.2 3% 4% 0% 7% 3% 0% 2.7 2.7 0.3 3.0 3.4 3.1 Red maple Assessment area 221E 221F M221A M221B M221C 16.0 13.8 23.6 18.0 17.5 15.5 30.0 30.1 40.0 25.2 26.8 30.9 88% 118% 69% 40% 53% 99% 30.5 30.4 40.3 27.5 27.0 31.3 2% 1% 1% 9% 1% 1% Red spruce Assessment area 221E 221F M221A M221B M221C 1.0 0.3 0.0 3.4 2.5 0.3 0.3 0.1 0.0 1.0 0.9 0.1 -67% -67% 0% -71% -64% -67% 0.3 0.1 0.0 1.0 0.9 0.1 0% 0% 0% 0% 0% 0% 2.5 1.8 1.6 4.7 4.8 1.1 -66% -62% 23% -63% -67% -84% 2.4 1.8 1.5 5.1 4.4 1.0 -2% 0% -6% 9% -8% -9% 1.3 0.7 0.4 1.9 3.9 0.5 -46% -61% -75% -60% -19% -55% 0.2 -68% 0.3 -63% 0.0 0% 0.1 -75% 0.0 -100% 0.0 0% 0.2 0.4 0.0 0.1 0.1 0.0 43% 33% 0% 0% NA 0% 0.3 0.5 0.0 0.2 0.1 0.0 81% 67% 0% 100% NA 0% 11.9 10.2 10.9 13.6 16.6 11.3 3.6 4.0 2.7 3.6 2.8 3.6 -70% -61% -75% -74% -83% -68% 6.0 5.2 9.4 5.6 7.2 5.3 68% 30% 248% 56% 157% 47% 5.2 3.8 9.0 5.6 7.3 4.5 45% -5% 233% 56% 161% 25% 1% 0% 0% 7% 3% -3% 7.3 6.9 0.7 10.3 9.8 7.6 5.4 5.4 1.1 4.8 7.1 5.9 -27% -22% 57% -53% -28% -22% 5.6 5.4 1.1 6.1 7.4 6.1 4% 0% 0% 27% 4% 3% 5.6 5.1 1.0 6.9 7.9 6.0 4% -6% -9% 44% 11% 2% 30.3 29.9 40.6 27.1 27.7 30.6 1% -1% 2% 8% 3% -1% 77.4 72.7 81.6 66.8 77.0 97.2 66.9 66.2 143.5 63.4 48.5 50.4 -13% -9% 76% -5% -37% -48% 70.3 68.7 145.9 75.9 50.6 52.8 5% 4% 2% 20% 4% 5% 74.5 70.6 160.7 81.8 56.6 55.1 11% 7% 12% 29% 17% 9% 0.3 0.1 0.0 1.0 0.9 0.1 0% 0% 0% 0% 0% 0% 5.1 1.7 0.0 17.0 13.0 1.4 0.4 -92% 0.0 -100% 0.0 0% 0.7 -96% 1.7 -87% 0.1 -93% 0.4 0.0 0.0 0.7 1.6 0.0 0% 0% 0% 0% -6% -100% 0.2 0.0 0.0 0.7 0.9 0.0 0% 0% 0% 0% -47% -100% (coninued on next page) APPENDiX 4 297 Tree species Current climate BA in 2009 Change BA in from 2100 2009b PCM B1 GFDL A1Fi Change from Change from BA in current BA in current 2100 climatec 2100 climatec Current climate PCM B1 GFDL A1Fi Change Change from Change from TPA in TPA in from TPA in current TPA in current 2009 2100 2009b 2100 climatec 2100 climatec Tree species Seciona Scarlet oak Assessment area 221E 221F M221A M221B M221C 4.8 3.4 0.3 6.7 7.4 7.8 0.7 0.4 0.1 1.0 1.2 1.1 -86% -88% -67% -85% -84% -86% 0.7 0.5 0.1 1.1 1.2 1.1 8% 25% 0% 10% 0% 0% 0.7 0.4 0.1 1.0 1.2 1.0 -2% 0% 0% 0% 0% -9% 4.0 2.5 0.5 6.4 6.7 6.2 1.1 0.7 0.3 1.5 2.0 1.8 -72% -72% -40% -77% -70% -71% 1.2 0.7 0.3 1.8 2.0 1.8 3% 0% 0% 20% 0% 0% 0.9 0.5 0.2 1.3 2.0 1.3 -17% -29% -33% -13% 0% -28% Sugar maple Assessment area 221E 221F M221A M221B M221C 6.1 5.9 7.0 6.8 6.4 5.5 8.0 8.0 8.3 9.2 8.0 6.8 30% 36% 19% 35% 25% 24% 8.0 8.0 8.3 9.5 8.0 6.7 0% 0% 0% 3% 0% -1% 7.3 7.2 7.7 8.0 7.8 6.1 -9% -10% -7% -13% -3% -10% 54.1 53.4 73.8 47.2 49.9 54.7 53.6 47.7 43.4 91.4 59.9 46.3 -1% -11% -41% 94% 20% -15% 50.8 44.9 42.6 88.0 57.3 43.0 -5% -6% -2% -4% -4% -7% 23.4 18.3 16.7 27.4 43.5 16.8 -56% -62% -62% -70% -27% -64% Tulip tree Assessment area 221E 221F M221A M221B M221C 6.5 6.6 2.1 5.8 5.7 9.9 10.5 11.1 7.0 4.6 9.6 15.5 62% 68% 233% -21% 68% 57% 10.8 11.5 6.9 5.1 9.7 15.8 3% 4% -1% 11% 1% 2% 10.8 11.3 7.1 5.6 9.9 15.6 3% 2% 1% 22% 3% 1% 22.9 25.0 9.5 14.1 16.9 36.6 104.3 356% 113.0 352% 123.3 1198% 19.3 37% 95.2 463% 130.7 257% 112.1 123.8 127.2 27.7 97.6 137.4 8% 10% 3% 44% 3% 5% 114.8 123.3 128.1 39.6 105.3 139.3 10% 9% 4% 105% 11% 7% White ash Assessment area 221E 221F M221A M221B M221C 2.5 2.7 4.8 1.9 1.7 1.7 2.9 3.2 4.6 2.1 2.4 2.2 18% 19% -4% 11% 41% 29% 3.0 3.4 4.7 2.1 2.4 2.2 4% 6% 2% 0% 0% 0% 2.9 3.2 4.5 2.1 2.5 2.2 0% 0% -2% 0% 4% 0% 11.9 14.0 19.5 8.1 8.0 8.0 7.3 6.4 25.1 1.4 7.2 4.0 -39% -54% 29% -83% -10% -50% 7.9 7.5 24.4 1.6 7.3 4.4 8% 17% -3% 14% 1% 10% 6.8 5.3 24.3 1.7 7.7 4.0 -7% -17% -3% 21% 7% 0% White oak Assessment area 221E 221F M221A M221B M221C 5.7 6.6 1.1 5.6 5.6 5.7 8.4 9.5 1.9 7.8 8.5 8.5 46% 44% 73% 39% 52% 49% 8.6 9.7 1.9 8.4 8.6 8.7 2% 2% 0% 8% 1% 2% 8.3 9.2 1.9 8.4 8.8 8.4 -1% -3% 0% 8% 4% -1% 12.8 15.0 2.9 12.8 12.3 12.2 57.7 64.3 20.8 50.0 60.6 58.7 349% 329% 617% 291% 393% 381% 60.9 68.1 22.2 56.1 61.7 61.8 6% 6% 7% 12% 2% 5% 60.8 65.8 21.5 61.0 65.1 61.7 5% 2% 3% 22% 7% 5% Assessment area values were derived from the weighted average of secions. Change under current climate represents the diference from 2009 through 2100 due to succession and management, but not climate. c Change from current climate for PCM B1 and GFDL A1FI represents the diference between these scenarios and current climate in 2100 and represents the potenial change due to climate change. a b APPENDiX 4 298 Table 42 (coninued). APPENDiX 4 Figure 61.—Projected changes in basal area for 17 species across the assessment area for PCM B1. Assessment area values were derived from the weighted average of secions. The width of the colored line represents the basal area for each species at various points through ime. For example, red maple had the highest basal area in 2010, and basal area is projected to increase by 2 percent for the PCM B1 scenario, in addiion to the projected 88-percent increase due to natural succession and management (see also Table 42). Figure 62.—Projected changes in trees per acre for 17 species across the assessment area for PCM B1. Assessment area values were derived from the weighted average of secions. The width of the colored line represents trees per acre for each species at various points through ime. For example, red maple had the highest trees per acre in 2010, and the number of trees per acre is projected to increase by 5 percent for the PCM B1 scenario, parially ofseing the projected 13-percent decrease due to natural succession and management (see also Table 42). 299 APPENDiX 4 Figure 63.—Projected changes in basal area for 17 species across the assessment area for GFDL A1FI. Assessment area values were derived from the weighted average of secions. The width of the colored line represents trees per acre for each species at various points through ime. For example, red maple had the highest basal area in 2010, and basal area is projected to increase by 1 percent for the GFDL A1FI scenario, in addiion to the projected 88-percent increase due to natural succession and management (see also Table 42). Figure 64.—Projected changes in trees per acre for 17 species across the assessment area for GFDL A1FI. Assessment area values were derived from the weighted average of secions. The width of the colored line represents trees per acre for each species at various points through ime. For example, red maple had the highest trees per acre in 2010, and the number of trees per acre is projected to increase by 11 percent for the GFDL A1FI scenario, almost ofseing the projected 13-percent decrease due to natural succession and management (see also Table 42). 300 APPENDiX 5: VuLNERABiLiTY AND CoNFiDENCE DETERMiNATioN EXPERT PANEL PRoCESS To assess vulnerabilities to climate change for each natural community type, we elicited input from a panel of 19 experts from a variety of land management and research organizations across the assessment area. We sought a team of panelists who would be able to contribute a diversity of subject area expertise, management history, and organizational perspectives. Most panelists had extensive knowledge about the ecology, management, and climate change impacts on forests in the assessment area. This panel was assembled at an in-person workshop in Morgantown, WV, in April 2013. Here we describe the structured discussion process that the panel used. Name Organizaion Jarel Barig Wayne Naional Forest Scot Bearer The Nature Conservancy - Pennsylvania Steve Blat Wayne Naional Forest Andrea Brandon The Nature Conservancy - West Virginia Patricia Butler* Michigan Technological University & Northern Insitute of Applied Climate Science Stephanie Connolly Monongahela Naional Forest Tim Culbreth Maryland Department of Natural Resources, Forest Service William Dijak U.S. Forest Service, Northern Research Staion Wade Dorsey Savage River State Forest Neil Gillies Cacapon Insitute Louis Iverson U.S. Forest Service, Northern Research Staion Maria Janowiak* U.S. Forest Service, Northern Research Staion & Northern Insitute of Applied Climate Science Kent Karriker Monongahela Naional Forest Dave Minney Independent consultant Coton Randall Ohio Department of Natural Resources, Division of Forestry Tom Schuler U.S. Forest Service, Fernow Experimental Forest Bill Stanley The Nature Conservancy - Ohio Al Steele U.S. Forest Service, Northeastern Area, State & Private Forestry Susan Stout U.S. Forest Service, Northern Research Staion Jason Teets Natural Resources Conservaion Service Frank Thompson U.S. Forest Service, Northern Research Staion *Workshop facilitator 301 APPENDiX 5 FoREST SYSTEMS ASSESSED The authors of this assessment modified and combined NatureServe (2011) ecological systems in order to describe specific forest ecosystems within the assessment area (see Chapter 1). For each forest ecosystem, we collected information related to the major system drivers, dominant species, and stressors that characterize that ecosystem from the relevant ecological literature. The panel was asked to comment on the forest ecosystem descriptions, and those comments were used to revise the descriptions in Chapter 1. PoTENTiAL iMPACTS To examine potential impacts, the panel was given several sources of background information on past and future climate change in the region (summarized in Chapters 3 and 4) and projected impacts on dominant tree species (summarized in Chapter 5). The panel was directed to focus on impacts to each forest ecosystem from the present through the end of the century, but more weight was given to the endof-century period. The panel assessed impacts by considering a range of climate futures bracketed by two scenarios: GFDL A1FI and PCM B1. Panelists were then led through a structured discussion process to consider this information for each forest ecosystem in the assessment. Potential impacts on ecosystem drivers and stressors were summarized based on climate model projections, the published literature, and insights from the panelists. Impacts on drivers were considered positive or negative if they would alter system drivers in a way that would be more or less favorable for that forest ecosystem. Impacts on stressors were considered negative if they increased the influence of that stressor or positive if they decreased the influence of that stressor on the forest ecosystem. Panelists were also asked to 302 consider the potential for climate change to facilitate new stressors in the assessment area over the next century. To assess potential impacts on dominant tree species, the panelists examined results from three forest impact models (Tree Atlas, LINKAGES, and LANDIS-PRO), and were asked to consider those results in addition to their knowledge of life history traits and ecology of those species. The panel evaluated how much agreement existed within the available information, between climate scenarios, and across space and time. Finally, panelists were asked to consider the potential for interactions among anticipated climate trends, species impacts, and stressors. Input on these future ecosystem interactions relied primarily on the panelists’ expertise and judgment because there are not many examples of published literature on complex interactions, nor are future interactions accurately represented by forest impact models (Box 12). ADAPTiVE CAPACiTY Panelists discussed the adaptive capacity of each forest ecosystem based on their ecological knowledge and management experience. Panelists were told to focus on characteristics that would increase or decrease the adaptive capacity of that system. Factors that the panel considered included characteristics of dominant species within each forest ecosystem (e.g., dispersal ability, genetic diversity, range limits) as well as comprehensive ecosystem characteristics (e.g., functional and species diversity, tolerance to a variety of disturbances, distribution across the landscape). The panelists were directed to base their considerations on the current condition of the system given past and current management regimes, with no consideration of potential adaptation actions that could take place in the future. APPENDiX 5 Box 12: A Note on Forest impact Models used in this Assessment During the expert panel workshop, preliminary LANDIS PRO results were used that included the climate parameters based on the average climate at the center of an ecological secion. This methodology was found to be less efecive in areas of complex topography, where steep elevaional gradients result in climaic gradients as well. The inclusion of climate data from the center of the secion projected unrealisic responses for several tree species in the Allegheny Mountain and Northern Ridge and Valley secions of the assessment area. The climate value at the center of the secion happened to fall on a highelevaion area, which represented the climate for the whole region as being much colder than the average climate. This was ideniied as an issue with the preliminary LANDIS PRO model results at the expert panel workshop, and many paricipants expressed their tendency to “discount” the results for black cherry and tulip tree. Following the expert panel workshop, the LANDIS PRO model was recalibrated with alternate climate data that beter represent the average climate. All results summarized in Chapters 5 and 6 were veted with the expert panelists to ensure their vulnerability rankings were sill consistent with the inal LANDIS PRO model results. VuLNERABiLiTY CoNFiDENCE After extensive group discussion, each panelist evaluated the potential impacts to and adaptive capacity of each forest ecosystem to arrive at a vulnerability rating. Participants were provided with individual worksheets and asked to list which impacts they felt were most important to that system in addition to the major factors that would contribute to the adaptive capacity of that ecosystem (Fig. 5). Panelists were also directed to give a confidence rating to each of their individual vulnerability determinations (Fig. b). Panelists were asked to evaluate the amount of evidence they felt was available to support their vulnerability determination and the level of agreement among the available evidence (Mastrandrea et al. 2010). Panelists evaluated confidence individually and as a group, in a similar fashion to the vulnerability determination. Panelists were directed to mark their rating in two-dimensional space on the individual worksheet and on a large group poster (Fig. a). This vulnerability figure required the participants to evaluate the degree of potential impacts related to climate change as well as the adaptive capacity of the ecosystem to tolerate those impacts (Swanston and Janowiak 2012). Individual ratings were compared and discussed and used to arrive at a group determination. In many cases, the group determination was at or near the centroid of all individual determinations. Sometimes the group determination deviated from the centroid because further discussion convinced some group members to alter their original response. Vulnerability and Conidence Figures For reference, figures of individual and group determinations for all nine forest ecosystems considered in this assessment are displayed in Figures 7 through 75. In each figure, individual panelist votes are indicated with a small circle and the group determination is indicated with a large square. We do not intend for direct comparison between these figures because the axes represent subjective, qualitative scales. 303 APPENDiX 5 Example Vulnerability Determinaion Worksheet Name: Ecosystem/Forest Type: How familiar are you with this ecosystem? (circle one) Low Medium high I have some basic knowledge about this system and how it operates I do some management or research in this system, or have read a lot about it. I regularly do management or research in this system What do you think are the greatest potenial impacts to the ecosystem? What factors do you think contribute most to the adapive capacity of the ecosystem? Vulnerability Determinaion Conidence Raing Use the handout for the vulnerability determinaion process and the notes that you have taken to plot your assessment of vulnerability on the igure below. Use the handout for the conidence raing process and the notes that you have taken to rate conidence using the igure below. The raings above are for the enire analysis area. Please note where you think potenial impacts or adapive capacity may vary substanially within the analysis area (e.g., forests in the eastern porion may be more prone to impact X). Figure 65.—Worksheet used for vulnerability and conidence determinaion by expert panelists, based on Swanston and Janowiak (2012). 304 APPENDiX 5 (a) (b) Figure 66.—Figure used for (a) vulnerability determinaion by expert panelists, based on Swanston and Janowiak (2012), and (b) conidence raing among expert panelists, adapted from Mastrandrea et al. (2010). Figure 67.—Vulnerability and conidence determinaions for Appalachian (hemlock)/northern hardwood forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. 305 APPENDiX 5 Figure 68.—Vulnerability and conidence determinaions for dry calcareous forest, woodland, and glade. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. Figure 69.—Vulnerability and conidence determinaions for dry oak and oak/pine forest and woodland. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. 30 APPENDiX 5 Figure 70.—Vulnerability and conidence determinaions for dry/mesic oak forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. Figure 71.—Vulnerability and conidence determinaions for large stream loodplain and riparian forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. 307 APPENDiX 5 Figure 72.—Vulnerability and conidence determinaions for mixed mesophyic and cove forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. Figure 73.—Vulnerability and conidence determinaions for north-central interior maple/beech forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. 308 APPENDiX 5 Figure 74.—Vulnerability and conidence determinaions for small stream riparian forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. Figure 75.—Vulnerability and conidence determinaions for spruce/ir forest. Circles indicate individual determinaions by each panelist and squares indicate the group determinaion ater consensus was reached. 309 APPENDiX 5 VuLNERABiLiTY STATEMENTS REFERENCES Recurring themes and patterns that transcended individual forest ecosystems were identified and developed into the vulnerability statements in boldface and supporting text in Chapter . The lead author developed the statements and supporting text based on workshop notes and literature pertinent to each statement. An initial confidence determination (evidence and agreement) was assigned based on the lead author’s interpretation of the amount of information available to support each statement and the extent to which the information agreed. Each statement and its supporting literature discussion were sent to the expert panel for review. Panelists were asked to review each statement for accuracy, whether the confidence determination should be raised or lowered, if there was additional literature that was overlooked, and if there were any additional statements that needed to be made. Any changes that were suggested by a single panelist were brought forth for discussion and approved by the entire panel. Mastrandrea, M.D.; Field, C.B.; Stocker, T.F.; Edenhofer, O.; Ebi, K.L.; Frame, D.J.; Held, H.; Kriegler, E.; Mach, K.J.; Matschoss, P.R.; Plattner, G.-K.; Yohe, G.W.; Zwiers, F.W. 2010. Guidance note for lead authors of the IPCC Fifth Assessment Report on consistent treatment of uncertainties. Geneva, Switzerland: Intergovernmental Panel on Climate Change (IPCC). Available at http://www.ipcc. ch/activities/activities.shtml. (Accessed Feb. 28, 2011). 310 NatureServe. 2011. NatureServe Explorer. Arlington, VA. Available at http://www. natureserve.org/explorer/classeco.htm. (Accessed September 11, 2014). Swanston, C.W.; Janowiak, M.K. 2012. Forest adaptation resources: climate change tools and approaches for land managers. Gen. Tech. Rep. NRS-87. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 121 p. Available at http://www.nrs.fs.fed. us/pubs/40543. (Accessed August 21, 2014). Butler, Patricia R.; Iverson, Louis; Thompson, Frank R., III; Brandt, Leslie; Handler, Stephen; Janowiak, Maria; Shannon, P. Danielle; Swanston, Chris; Karriker, Kent; Bartig, Jarel; Connolly, Stephanie; Dijak, William; Bearer, Scott; Blatt, Steve; Brandon, Andrea; Byers, Elizabeth; Coon, Cheryl; Culbreth, Tim; Daly, Jad; Dorsey, Wade; Ede, David; Euler, Chris; Gillies, Neil; Hix, David M.; Johnson, Catherine; Lyte, Latasha; Matthews, Stephen; McCarthy, Dawn; Minney, Dave; Murphy, Daniel; O’Dea, Claire; Orwan, Rachel; Peters, Matthew; Prasad, Anantha; Randall, Cotton; Reed, Jason; Sandeno, Cynthia; Schuler, Tom; Sneddon, Lesley; Stanley, Bill; Steele, Al; Stout, Susan; Swaty, Randy; Teets, Jason; Tomon, Tim; Vanderhorst, Jim; Whatley, John; Zegre, Nicholas. 2015. Central Appalachians forest ecosystem vulnerability assessment and synthesis: a report from the Central Appalachians Climate Change Response Framework project. Gen. Tech. Rep. NRS-146. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research Station. 310 p. Forest ecosystems in the Central Appalachians will be affected directly and indirectly by a changing climate over the 21st century. This assessment evaluates the vulnerability of forest ecosystems in the Central Appalachian Broadleaf ForestConiferous Forest-Meadow and Eastern Broadleaf Forest Provinces of Ohio, West Virginia, and Maryland for a range of future climates. Information on current forest conditions, observed climate trends, projected climate changes, and impacts on forest ecosystems was considered by a multidisciplinary panel of scientists, land managers, and academics in order to assess ecosystem vulnerability to climate change. Appalachian (hemlock)/northern hardwood forests, large stream floodplain and riparian forests, small stream riparian forests, and spruce/fir forests were determined to be the most vulnerable. Dry/mesic oak forests and dry oak and oak/pine forests and woodlands were determined to be least vulnerable. Projected changes in climate and the associated impacts and vulnerabilities will have important implications for economically valuable timber species, forest-dependent wildlife and plants, recreation, and long-term natural resource planning. KEy WORDS: climate change, vulnerability, adaptive capacity, forests, Climate Change Tree Atlas, DISTRIB, LANDIS PRO, LINKAGES, expert elicitation, climate projections, impacts The U.S. Department of Agriculture (USDA) prohibits discrimination in all its programs and activities on the basis of race, color, national origin, age, disability, and where applicable, sex, marital status, familial status, parental status, religion, sexual orientation, genetic information, political beliefs, reprisal, or because all or part of an individual’s income is derived from any public assistance program. (Not all prohibited bases apply to all programs.) Persons with disabilities who require alternate means for communication of program information (Braille, large print, audiotape, etc.) should contact USDA’s TARGET Center at 202-720-2600 (voice and TDD). To file a complaint of discrimination, write to USDA, Director, Office of Civil Rights, 1400 Independence Avenue, S.W., Washington, DC 20250-9410, or call 800-795-3272 (voice) or 202-720-6382 (TDD). USDA is an equal opportunity provider and employer. Printed on recycled paper Northern Research Station www.nrs.fs.fed.us