PROCEEDINGS of the Third Longleaf Alliance Regional Conference

FOREST FOR OUR FUTURE Restoration and Management of Longleaf Pine Ecosystems: Silvicultural, Ecological, Social, Political and Economic Challenges

Hosted by The Longleaf Alliance, USDA Forest Service, and US Geological Survey Alexandria, Louisiana October 16-18, 2000

Longleaf Alliance Report No. 5 July 2001 THE HISTORY AND CHARACTER OF LONGLEAF PINE SYSTEMS IN THE WEST

Latimore Smith (The Nature Conservancy – Louisiana Field Office Baton Rouge, LA)

Longleaf pine (Pinus palustris) historically ranged from southern Virginia to east Texas, occupying mainly the sandy and silty Atlantic and Gulf Coastal Plains. In the West Gulf Coastal Plain (WGCP), longleaf pine historically occupied over 7 million acres in west Louisiana and southeast Texas, occurring on a wide variety of geologic substrates, and soil conditions, from high hills to flatwoods. It was principally found on acidic sandy loams and silt loams formed on Pleistocene-aged deposits to the south, but also on acidic clays, loams and sands of older Tertiary formations to the north. Such variation in substrates supporting longleaf pine produced a truly amazing complexity of herbaceous ground cover communities in the longleaf system of the WGCP, and notable variation in the structure of the virgin forests and savannas.

The western range of longleaf pine of longleaf pine, west of central Mississippi, is comprised entirely of what has been generally labeled the “longleaf-bluestem” range, due to the dominance of bluestems and broomsedges (mainly little bluestem, Schizachyrium scoparium) and complete lack of wiregrass (Aristida berychiana/stricta) in the ground cover. While much emphasis has been placed on the longleaf-wiregrass system of the eastern states by many in recent years, the longleaf-bluestem range is equally worthy of attention. It supports a grand assortment of native species, both plant and (many rare and restricted to longleaf communities), and has been equally profoundly reduced in extent by human endeavor.

As throughout the range of longleaf, frequent fire was, without question, the primary force that gave rise to and maintained the longleaf system in the western part of its range. Fires originating from lightning (long before the arrival of aboriginal peoples, e.g., Komarek 1968) are estimated to have burned through the grassy ground cover of the western forests and savannas every 1 to 5 years, mainly in the spring and early summer. There are clear indications in early written accounts (e.g., Sargent 1884, Mohr 1897) that the virgin longleaf forests of the western range were typically more heavily stocked than the majority of the forests to the east, almost certainly owing to the predominance of more fertile loams to the west. Well-drained sites in the loamy flatwoods and gently rolling hills are particularly conducive to excellent growth of longleaf pine. Despite such heavy average stocking of trees, the forest was naturally heterogeneous in stand structure, as it was to the east, typically varying over short distances from very densely stocked areas to very open areas with few, if any, trees (e.g., Schwarz 1907). The virgin forests of the west were also primarily uneven-aged, individual stands usually comprised of trees of many ages and sizes (e.g., Chapman 1909), although more-or- less even-aged stands did originate after devastating windstorms. The largest longleaf pines in the virgin forests of the west were 35 to 40 inches in diameter breast height (dbh) and about 120 feet tall, with the oldest trees averaging 200 – 300 years of age, but rarely exceeding 400 years (Mohr 1897).

The two principal historic landscape settings for longleaf pine in its western range were forests (or woodlands) in rolling hills, and pine savannas and flatwoods in the flats. Although there were a variety of landscape and substrate settings on which longleaf was found in the hills, longleaf was most common on dry- mesic, acidic sandy loams, with essentially pure forests of longleaf pine stretching across the rolling hills. The most common hardwoods (yet very scattered, in general) found in these forests included southern red oak (Quercus falcata), post oak (Q. stellata), blackjack oak (Q. marilandica), mockernut hickory (Carya alba), and black hickory (C. texana). In this landscape, longleaf characteristically transitioned quickly going down major slopes to a mixed shortleaf pine (Pinus echinata)-hardwood type that formed narrow belts along upper reaches of major slopes leading to permanent streams. Down slope from the latter type was typically a mesic mixed loblolly pine (P. taeda)-hardwood type. Judging from best remaining examples, the ground cover plant communities of the dry-mesic, loamy longleaf uplands was extremely diverse. Field studies of best remnants of the type reveal that over 100 herbaceous plant species are often present on a typical acre. While grasses (particularly little bluestem), composits (sunflower family) and legumes are the most common groups, a huge assortment of other species are present. Many of these species are restricted to upland longleaf systems, or reach their greatest abundance there. It is fair to estimate that over 400 plant species are present in the loamy upland longleaf forests of the western range, and over 100 are essentially restricted to these forests in the region.

1 Other significant longleaf types in the rolling hills of the western longleaf range included xeric longleaf woodlands on deep sands, and xeric to mesic longleaf woodlands on acid clays. Xeric longleaf woodlands of the WGCP support a significant component of scrub oaks, such as bluejack oak (Quercus incana), sand post oak (Q. margarettiae), and blackjack oak, and probably burned less frequently than the typical dry-mesic longleaf forest of the loamy uplands due to sparser grassy fuels. Although overall species richness is not as great in the deep sandy uplands as the loamy uplands, numerous specialized herbaceous species are restricted to the sandy xeric uplands of the WGCP longleaf range such as scarlet catch-fly (Silene subciliata) and Winkler’s Indian blanket (Gaillardia aestivalis var. winkleri). Analogous sandhill longleaf sites in much of the East Gulf Coastal Plain support Gopher tortoise (Gopherus polyphemus), but this animal does not range west of the central Florida Parishes of southeastern Louisiana.

Sandstone glades supporting scattered, stunted longleaf pine and scrub oaks and a specialized flora are locally common (yet rare overall) in portions of the western longleaf range. Such glades arise on outcrops of sandstone “pavement”, mainly on the Catahoula formation in central and western Louisiana.

Inclusional areas of calcareous clays within the longleaf range of the WGCP supported mixed forest types (lacking longleaf), and calcareous prairie grasslands. In transitional areas between acid loams or sands with longleaf and calcareous clays supporting other vegetation types, there was typically a fairly dramatic transition in types over short distances. At the edges of the longleaf range in the west, pure longleaf forests often gave way to mixed forests of longleaf, shortleaf, loblolly and a variety of hardwoods. While transitions to other types were typically abrupt and transitional mixed forests including longleaf pine were never extensive in Louisiana, such forests apparently were prominent at the western edge of the longleaf range in Texas (Paul Harcombe, pers. comm.).

Hillside seepage bogs are embedded within frequently burned, sandy upland longleaf forests of the western range, being particularly abundant in west central Louisiana. This region may support the highest density of hillside seepage bogs found anywhere in the southeastern United States. In the heart of this “bog country”, dozens of seepage bogs are present per square mile. While hundreds of bogs are known from this area, each is usually less than 2 acres in size (though much larger ones exist), and the total acreage of hillside seepage bogs in the WGCP probably does not exceed 3,000 acres. These special wetlands are floristically closely related to the acidic, wet longleaf pine flatwood savannas of the region, sharing many species in common, but differ markedly from savannas in the WGCP by the presence and sheer abundance of insectivorous yellow pitcher- (Sarracenia alata; the only species in the genus found in the WGCP). These fascinating plants are completely and inexplicably absent from the flatwood savannas of the WGCP. Hillside seepage bogs support at least one regional endemic, Bog brown-eyed Susan (Rudbeckia scabrifolia).

Longleaf pine flatwood savannas dominated the youngest and flattest Pleistocene terraces of the outer coastal plain of western Louisiana and east Texas (north of the coastal prairie zone), and were found further inland on “relict” flats on higher parts of older Pleistocene terraces and certain Tertiary formations. Soils of the flatwoods of the WGCP are overwhelmingly acidic silt loam alfisols, although a very rare and unusual savanna type is found on sodic silt loams. Flatwood savannas are actually closely intermingled wetlands (generally called “savannas”) and non-wetlands (generally called “flatwoods”), produced by close-proximity variations in topography that greatly influence microsite hydrology. Wet pine savannas and non-wet flatwoods in the WGCP are tightly co-mingled in a landscape studded with “pimple mounds” (also called mima mounds), which are small, rounded knolls or hillocks, usually a few feet higher than the intermound areas, and that are more or less regularly distributed across the flats. In this unique landscape, “savannas” occupy the wet intermound areas and “flatwoods” occupy the well-drained pimple mounds, but the entire complex is usually referred to as a savanna. The federally-listed (“Endangered”) plant, American chaffseed (Schwalbea americana), was recently rediscovered in southwest Louisiana, after a near 100-year hiatus, growing on pimple mounds in the longleaf pine flatwoods. The flatwood savannas of the west, as those in the east, are very rich in native herbaceous species, dominated by grasses and sedges, but a multitude of other groups are represented. The close intermingling of wet and non-wet positions produces very high species richness on an average acre in the western savannas. Herb-dominated “flatwood ponds” occupy the alluviated remnants of ancient stream beds that wind throughout the WGCP flatwoods.

2 Today, little more than 5 percent of the longleaf acreage present 120 years ago in the WGCP remains relatively intact. While the majority of the remaining acreage is present on federal land (Kisatchie National Forest, National Forests of Texas, Fort Polk Military Reservation), significant holdings are present on private land. In particular, The Nature Conservancy (TNC) has acquired ecologically important areas in the lower flatwoods of southwest Louisiana, and other longleaf areas in southeast Texas. Future plans by TNC specify much emphasis on the conservation and restoration of longleaf systems in the west.

Literature Cited Chapman, H. H. 1909. A method of studying growth and yield of longleaf pine applied in Tyler Co., Texas. Soc. Am. For. Proc. 4:207-220. Komarek, E.V. 1968. Lightning and lightning fires as ecological forces. Proc. Ann. Tall Timbers Fire Ecology Conf., Tallahassee, FL. 8:169-197. Mohr, C. 1897. The timber pines of the southern United States. Government Printing Office, Washington. Sargent, C. S. 1884. Report on the forests of North America (exclusive of Mexico). U.S. Dept. Int. Census Off. 10th Census Rept., vol. 9, Washington, D.C. Schwarz, G. F. 1907. The longleaf pine in virgin forest. John Wiley and Sons, New

3 ECONOMIC AND MARKETING CONSIDERATIONS FOR ENHANCING LONGLEAF PINE MANAGEMENT BY PRIVATE FOREST LANDOWNERS

Michael A. Dunn (Extension Natural Resources Program, Louisiana State University Agricultural Center)

INTRODUCTION Good morning. I want to spend a little time this morning discussing with you some opportunities for enhancing longleaf pine management. And I would like to do this from the point of view of not only an economist, but also an extension specialist. I am both. The economics training provides me with theory, ideas, and a way to think about how things should work. My extension experience tempers that with practicality, with what works in the “real world.” My goals with this presentation are to 1) examine the critical economics and marketing issues that need to be addressed if longleaf ecosystems are to be restored on private lands, and 2) to address practical considerations that must be addressed if longleaf areas are to flourish on private lands.

THE POWER OF INFORMATION When we talk about the market place, we need to understand that one of the most critical needs in the marketplace is the need for information. All one has to do is witness the explosion in information technology over the last twenty years to understand its importance. Therefore, information can be our greatest asset. Conversely, misinformation can lead to poor choices. This has often been the case in forested systems. We have very often fallen victim to bad information. It is not necessarily due to negligence but rather to the complex systems we try to manage and for which we seek information.

This lack of good information, or consumption of bad information, has made the forestry community a skeptical one. I do not think this is a bad thing. Forestry is not an insignificant investment. It is a long-term commitment to manage a multitude of resources. Quite often, many years pass before the manager or landowner sees the results of her efforts. Therefore, information becomes an even more important commodity in natural resources, particular in forest ecosystem management where profit forms part of the utility function.

CRITICAL KEYS FOR SUCCESS I believe that longleaf can be a valuable ecosystem, even from the hardened economist’s perspective. More and more, the evidence mounts that it can be financially competitive with some of our other species, provided it is put on the proper sites and provided it is managed correctly. The truth is, we still lack enough good scientific information on a broad scale to make a determination regarding the financial viability of longleaf.

But even if it cannot compete financially with loblolly pine, does this mean that we must discard it as a preferred species? Remember, economics does not necessarily equate to dollars and cents. Economics looks at human behavior in the face of scarcity and endeavors to map or predict this behavior. Economists look at many other things besides merely whether or not something is profitable. In the case of natural resource management, many people do not count maximizing profit as their top priority, or even a priority at all. For some, there are other motivations. Economists are just as interested in these non-monetary motivations.

Regardless of the motivation, however, there has to be a strategy implemented by the proponents of longleaf pine ecosystem restoration in order to see positive results. I see the following as critical keys for success in longleaf restoration: 1. Good economic information 2. A thorough understanding of the target group 3. A good marketing strategy 4. Good programs to promote longleaf management 5. Persistence, patience, and constant re-evaluation 6. Building a good support system

In order to start our understanding of our longleaf ecosystem and its many benefits, we are going to have to learn more about its economic behavior. This means we need to learn more about its profit potential, its long-term viability as a crop species, and its non-monetary benefits. We can’t just think it’s a good species;

4 we have to know it’s a good species, know ALL the reasons why it’s a good species, and be able to back that up with data. When we are able to sufficiently capture this information, we can truly begin the process of promoting longleaf.

The marketing process actually begins with performing some analysis. Do we really know the target group that might be interested in growing longleaf? Have we made the proper identification, or are we spending a lot of time trying to persuade a group of people that are very risk averse and will always view longleaf pine management as risky, whether or not it is warranted. Once we have this information, we can then begin establishing a strategy aimed at the proper audience to bring about our desired outcomes. This includes developing some good programs to promote longleaf ecosystem restoration and management. The extension specialist in me would like to think that the Cooperative Extension System could play a big role in the implementation of these programs. However, an extension specialist or agent is only as good as the research to which he’s privy. Therefore, research and extension must work hand in hand as a positively reinforcing system. Extension must assist research with feedback regarding what sorts of research would be helpful to the longleaf pine management community, and research must provide extension with relevant, applied data. Once this sort of system is in place, we can begin the process of truly developing and implementing a program to re-introduce longleaf pine in areas where it has not existed for a long time.

THE OLD PARADIGM We have all heard the following statements before: “Longleaf can’t compete with loblolly.” “Longleaf stays in the grass stage too long.” “Longleaf is too expensive to manage compared to loblolly.” “It’s too much trouble.” “It takes too long to recover investment costs because of the longer rotation.” “It’s too risky, because you have to use fire and the general public doesn’t like us burning all the time.”

If we as resource managers are ever to be successful at really developing a viable longleaf ecosystem, we must address all of the above concerns, the remnants of “the old paradigm.” Perhaps we can even develop or create “a new paradigm” if we can answer the following questions: 1) Can longleaf compete “head to head” with loblolly? 2) Can we get it out of the grass stage quickly? 3) Can we manage it on a cost basis competitive with loblolly? 4) Can we, through enhanced genetics and management techniques, shorten its overall optimal financial rotation? 5) Can we lower the risks associated with longleaf, risks such as those inherent with using frequent fire and providing habitat for endangered species such as the red-cockaded woodpecker? By tackling these questions and developing research and information that addresses these issues, we can perhaps affect a small shift in “the old paradigm.”

SHOW ME THE MONEY Many NIPF’s do manage for financial gain. For them, it is essential that we provide them with good accurate information regarding the financial potential of longleaf. Nothing less will do. It is essential to 1) provide them with studies that show longleaf is competitive, 2) continue to develop new management techniques that improve the competitiveness of longleaf, and 3) show them through demonstration projects that it really can work. Don’t just tell them; show them. Landowners like to see it happening in real time, not just some abstraction presented as a slide show or talked about in a classroom or lecture hall. We have had great success in extension using demonstration projects. More demonstration projects for longleaf ecosystems would be very beneficial.

Are these studies and demonstration projects, in and of themselves, enough? I don’t think so. Leadership among NIPF’s tend to be in close communication with professional foresters. Professional foresters, as a group, tend to be very skeptical and conservative. They tend to be industry oriented; therefore, they tend to be loblolly pine oriented. As a group, it is very difficult to change their minds once they have decided on something, like how to manage or what to manage. Therefore, longleaf pine proponents are going to have to target this group for education as well. Therefore, they need to accomplish tasks such as developing widely available growth and yield models that are “data defensible” and practical. These models should be user friendly, or computerized in a user friendly, intuitive format. We need to train foresters on the use of good

5 growth and yield models. And we need more information from the experts regarding the pros and cons of various growth and yield models.

We also need good programs to promote sustainable management of a variety of forest ecosystem types, not just loblolly or longleaf, for that matter, but a more diverse overall forest ecosystem. In Louisiana, we have many outstanding programs that promote active forest management and management for more diverse ecosystems, such as the Forest Stewardship Program and the Forest Productivity Program. These programs provide subsidies of one sort or another, whether it is cash reimbursement for part of one’s silvicultural costs or assistance with writing management plans. As an economist, I am no big fan of subsidies. However, the extension specialist in me has seen the benefits of demonstration and information access. By having programs like this, it allows for more adoption of management of diverse forest ecosystems. As more and more landowners and forest managers see these management regimes working, more and more will adopt these strategies. And then, hopefully, we will no longer need the subsidies.

TWO MARKETING STRATEGIES I would like to offer to you today two marketing strategies that I believe might assist in developing adoption of longleaf pine as a viable alternative to other species. The first is an informational campaign. If longleaf proponents can collect or conduct quality, applied research that can be put to use “on the ground,” this will go a long way in starting the process of increasing longleaf’s viability. However, the results of this research must be successfully disseminated to and then through the proper outlets for the information to be useful. It is not enough to pass the research to extension specialists and then leave it at that, hoping the specialists will incorporate it into her daily educational regime. No, extension specialists must be convinced as well. Therefore, it is important for longleaf proponents to be patient and to engage in long-term information adoption strategies. People do not change their minds overnight and the more it is forced on them the more they will resist. They must be skillfully and subtly led to the information. They must see the information over and over again, in different formats. There are very few “instant hits” in this world. If there were, we would only see those annoying commercials on television once. Repetition is very important. Once key personnel in leadership roles can be recruited as “advocates,” then it becomes easier to promote adaptation of longleaf pine management. However, this can only be accomplished through patience and persistence. Also, more demonstration areas should be established that would be used to educate and inspire private forest landowners as well as professional foresters.

The second strategy involves targeting what I call the “New and Emerging” private forest landowners. Longleaf proponents should remember that, especially today, not everyone who owns a forest wants to maximize financial profit from it. They may have many other motivations. These new and emerging landowners have a “new rural” attitude. They tend to be people that are fleeing the cities for a more rustic setting; however, they bring a more urban mentality into the countryside. One of their non-monetary objectives may be to provide good wildlife habitat or to manage for aesthetics. These people may be more receptive to managing for longleaf if it is marketed in a way that appeals to them. The trick is to learn what works.

CONCLUSIONS The most critical aspect of the economic decision making process is the collection of information. The better the information, the better the decisions.

Some traditional NIPF’s will invest in longleaf pine management if they can see a way to make it pay for them. It is up to us to provide them with that information, if it truly exists and as long as it is truthful and accurate. This information should include new innovations in silviculture or genetics that renders longleaf more financially competitive; new federal, state, or local policies and/or programs that reduce the risk associated with managing longleaf (rules, regulations, procedures, and cost share programs); and more information regarding management for alternative products that increase the net present value of longleaf ecosystem investments.

Some NIPF’s, especially the “new and emerging” private forest landowners that move further and further away from cities and into the country, derive maximum utility from their forestland from sources other than maximum profit. For these groups, we should strive to develop materials, conferences, workshops,

6 demonstration plots, and projects that show the many diverse benefits that result from the longleaf ecosystem. We should also find ways to pull landowners together in the longleaf range and develop incentives for these people to work to develop longleaf ecosystems that span property boundaries. I believe the proper financial incentives can be found to achieve this. We should remember that these new and emerging landowners might not be reached in the traditional way. They do not think or behave in traditional NIPF fashion; therefore, we should not treat them as traditional NIPF’s. Groups such as extension services and state forestry agencies and associations that have tended to cater to the traditional loblolly pine forest landowner should strive to develop partnerships with other stakeholder groups that desire to increase forestland diversity. We can all accomplish more by working together than we can by working individually.

Natural resource management professionals, particularly professional consulting foresters and landowner assistance foresters, need access to the best available technical information. Their informational and educational needs include a better understanding of growth and yield models, more information regarding alternative management scenarios which can yield intermittent or annual incomes from longleaf forests that have the effect of increasing profitability, more technical information that can assist them in managing longleaf for greater profit and reduced risk, and continuing education workshops designed to provide them with much of this information.

I believe that longleaf can be a viable financial alternative if we market its benefits and provide the proper groups with enough significant information. The Longleaf Alliance is certainly doing its part in promoting the benefits of the longleaf pine ecosystem. Through patience and persistence, I am confident that future generations will witness the revitalization of longleaf pine.

7 ARTIFICIAL REGENERATION: AN ESSENTIAL COMPONENT OF LONGLEAF PINE ECOSYSTEM RESTORATION

James P. Barnett (USDA Forest Service, Southern Research Station, 2500 Shreveport Highway, Pineville, LA 71360)

ABSTRACT: Regenerating longleaf pine by artificial means is an essential component of restoring the ecosystem across most of its range because there are limited acres of longleaf stands remaining. Establishing longleaf pine is an early step in the ecosystem restoration process. An overview discussion of artificial regeneration techniques and related issues are presented this paper. Other papers follow in the proceedings that provide more specific information related to reestablishing longleaf pine forests.

INTRODUCTION Although longleaf pine ecosystems once occupied over 90 million acres in the southeastern United States’ coastal plain from southern Virginia to central Florida and west to eastern Texas (Frost 1993), less than 3 million acres remain (Kelly and Bechtold 1990). Utilization of trees in the original forest was so complete when lumbering reached the western Gulf Coast Region that inadequate numbers of seed trees remained to naturally regenerate many of the harvested stands. So, artificial regeneration must be used to restore longleaf on the appropriate sites where it originally grew. Research indicates that productivity of an ecosystem is controlled to an overwhelming extent by the functional characteristics of the dominant plants (Grime 1997). So, with reestablishment and appropriate management, restoration processes will begin. Until recently, regeneration success from planting was generally unacceptable due to problems related to severe competing vegetation, delayed stem elongation, and poor storability of bare root seedlings.

Direct seeding of longleaf pine was widely used in the 1950’s and 1960’s to reforest large areas of cutover forests (Derr and Mann 1971). But, difficulty in controlling the density of stand stocking reduced the use of direct seeding. We now have the knowledge and technology to reestablish longleaf by planting bare rootstock, but meeting of the criteria to assure success is difficult to achieve on an operational basis. As a result, the use of container stock has increased dramatically. Now, about 50 million container longleaf seedlings are produced annually across the South. The quantity of container stock continues to increase and is limited primarily by the availability of good quality seeds.

A short discussion of reforestation related issues important in reestablishing longleaf pine forests follow and more detailed papers on these topics occur in the proceedings.

LONGLEAF PINE SEEDLING PRODUCTION Key to the effective production of longleaf pine nursery stock is availability of quality seeds. Longleaf pine seed collection, processing, storage, and treatment require exceptional care to maintain high quality (Barnett and Pesacreta 1993). Longleaf seed coats carry a significant population of pathogenic fungi and benefit from treatments that minimize disease losses (Barnett et al. 1999).

The knowledge and technology to successfully plant bareroot nursery stock have improved significantly in the last decade. The components of successful regeneration include: (1) well-prepared, competition-free sites; (2) healthy, top-quality, fresh planting stock; (3) meticulous care of stock from lifting to planting; (4) precision planting; and (5) proper post-planting care. Because controlling all five elements is difficult in most cases, planting success with bareroot stock remains elusive.

Numerous studies have demonstrated that under adverse planting conditions, such as poor sites, conditions of moisture stress, and out-of-season planting, container seedlings survive and grow better than bareroot stock (Barnett and McGilvray 1993, Boyer 1989). Guidelines for successfully producing longleaf container stock are now available (Barnett and McGilvray 1997).

PLANTING METHODS Successful establishment of longleaf pine seedlings requires that bareroot seedlings are lifted and planted promptly. Storage for more than a few days, even if refrigerated, results in poor survival. Treatment of seedling roots with a benomyl fungicidal dip does improve establishment if planting is delayed (Barnett et al.

8 1988). On the other hand, container stock can be held either in the containers or extracted and cold stored with good success.

The key to a seedling’s survival after planting is the ability of the root system to quickly begin taking up water and nutrients. Depending upon the site, either hand or machine planting can be the most efficient and reliable option. Large, level, open tracts are most efficiently planted by machine, while smaller tracts or sites with minimal site preparation are more easily hand-planted. Proper planting depth is critical for longleaf pine. Plant longleaf seedlings so the bud is not buried nor the root collar exposed.

SITE PREPARATION AND RELEASE Longleaf pine has long had a reputation as a difficult to regenerate species. Because longleaf seedlings have no initial stem growth, they are very sensitive to competition. Cultural treatments should be aimed at improving longleaf survival during the critical first year after planting by reducing competition. Initial site preparation that has as extended effect is highly desirable. It may not be necessary to apply post-planting competition control measures if the initial treatment is long lasting (Boyer 1989).

On many sites, post-planting release with a herbicide is an effective treatment for promoting rapid early growth of planted longleaf pine. Release is most important for seedlings with low levels of site preparation. Post-planting applications will likely be economical because early growth response may reduce the grass- stage by 1 to 2 years.

SEED MOVEMENT AND GENETICS The genetic variability of longleaf pine is less than for loblolly or slash pine. It is most important to avoid seed sources from the southern extremity of the range (Lantz and Kraus 1987). However, Schmidtling and Sluder (1995) found no consistent seed source differences in performance among physiographic provinces.

Considerable research has been conducted on evaluating regional seed sources for susceptibility to brown- spot needle blight and brown-spot resistance has been one factor incorporated into development of tree improvement programs.

COST-SHARE AND INCENTIVE PROGRAMS Although government funding for cost share programs to encourage reforestation has been decreasing, there are still incentives for planting longleaf pine seedlings (Karrfalt and Lantz 1998). Planting can be justified for programs other than timber production, e.g., wildlife and restoration. In addition, longleaf pine has a favored status in that there is a higher level of support permitted for longleaf that some other species.

Because of the harvests of pine exceeds growth in some areas of the South, a number of southern states are considering creating incentives for reforestation of small-landowner properties. An example of this is in Louisiana. Concern about sustaining the productivity of Louisiana’s forest lead the legislature to establish in 1997 the Louisiana Forest Productivity Program. The act directs that 75 percent of the state’s share from timber severance tax be used us support cost sharing for site preparation, planting or seeding, and control competing vegetation (Barnett 1999).

WEST GULF TREE IMPROVEMENT PROGRAM AND SEED ORCHARDS The West Gulf Tree Improvement Program is one of the premier programs with the objective of developing and producing genetic material in the western Gulf Region. This program--and the tree improvement efforts in the Southeast--are responsible for forest industry and state organizations being able to produce genetically improved seeds for loblolly and slash pines in quantities to meet their needs. Much less effort has been put on improving longleaf pines, but the West Gulf Tree Improvement Program does have an emphasis on longleaf pine. A number of longleaf pine seed orchards are in production.

The current seed orchards do not produce sufficient quantities of longleaf seeds to meet the demand, but improved seeds are becoming available in increasing quantities.

LITERATURE CITED Barnett, Jim. 1999. Tree planting—forests made by hand. Forests and People 49(2): 12-14.

9 Barnett, J.P.; Brissette, J.C.; Kais, A.G.; Jones, J.P. 1988. Improving field performance of southern pine seedlings by treating with fungicides before storage. Southern Jour. Applied Forestry 12: 281-285. Barnett, J.P.; McGilvray, J.M. 1993. Performance of container and bareroot loblolly pine seedlings on bottomlands in South Carolina. Southern Jour. Applied Forestry 17: 80-83. Barnett, J.P.; McGilvray, J.M. 1997. Practical guidelines for producing longleaf pine seedlings in containers. Gen. Tech. Rep. SRS-14. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 28 p. Barnett, J.P.; Pesacreta, T.C. 1993. Handling longleaf pine seeds for optimal nursery performance. Southern Jour. Applied Forestry 17: 180-187. Boyer, William D. 1989. Response of planted longleaf pine bare-root and container stock to site preparation and release: fifth-year results. In: Miller, J.H., compl. Proceedings of the fifth biennial southern silvicultural research conference. Gen. Tech. Rep. SO-74. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station: 165-168. Derr, H.J.: Mann, W.F., Jr. 1971. Direct-seeding pines in the South. Agric. Handbook 391. Washington, DC: U.S. Department of Agriculture, Forest Service. 68 p. Frost, C.C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. In: Proceedings of the Tall Timbers Fire Ecology Conference 18. Tallahassee, FL: Tall Timbers. Grime, J.P. 1997. Biodiversity and ecosystem function: the debate deepens. Science 227(5330): 1260-1261. Karrfalt, Robert P.; Lantz, Clark W. 1998. Current reforestation demands on southern nurseries. In: Landis, T.D.; Barnett, J.P., coord. National proceedings: forest and conservation nursery associations—1998. Gen. Tech. Rep. SRS-25. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station: 5-7. Kelly, J.F.; Bechtold, W.A. 1990. The longleaf pine resource. In: Proceedings of the symposium on the management of longleaf pine, ed. Farrar, R.M., Jr. Gen. Tech. Rep. SO-75. New Orleans, LA: USDA Forest Service, Southern Forest Experiment Station. Lantz, Clark W.; Kraus, John F. 1987. A guide to southern pine seed sources. Gen. Tech. Rep. SE-43. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station. 34 p. Schmidtling, R.C.; Sluder, E.R. 1995. Seed transfer and genecology in longleaf pine. In: Weir, R.J., coord. Proceedings 23th southern forest tree improvement conference; 1995 June 19-22; Asheville, NC. Raleigh, NC: Southern Forest Tree Improvement Committee: 78-85.

10 AN OVERVIEW OF LONGLEAF PINE ECOSYSTEMS OF THE WEST GULF COASTAL PLAIN

D. Craig Rudolph (US Forest Service, Southern Research Station, Nacogdoches, Texas)

Longleaf pine was once the dominant canopy species on 7-9 million acres of the West Gulf Coastal Plain. Logging activities, especially with the use of railroads, harvested nearly all of the original longleaf timber by 1930 (Maxwell and Baker 1983). Due to the combined effects of lack of seed source, conversion to other uses, and abundant livestock (especially feral hogs), regeneration was minimal (Frost 1993). Consequently, longleaf pine remains much diminished on the landscape to the present day.

Longleaf pine attained its present distribution through expansion from Pleistocene refugia in Florida or south Texas, based on fossil pollen profiles, during the last 10,000 years (Webb 1986). This expansion was possible due to the frequent fire regime resulting from ignitions due to lightning and Native Americans. Longleaf pine communities are maintained by frequent, low-intensity fires with characteristic return intervals of several times per decade.

Longleaf pine is able to maintain dominance in forest communities subject to frequent fire and outbreaks of southern pine due to a number of adaptations. The ability to produce viable seedlings and saplings in a frequent fire regime is possible due to large seed size, presence of a grass stage, thick clusters of needles that protect terminal buds, large diameter terminal buds, and rapid height growth out of the grass stage (Wahlenberg 1960, Platt et al. 1988). Resistance to southern pine beetles is due primarily to abundant resin production that hinders initial attack (Hodges et al. 1979, Conner and Rudolph 1995). Consequently, longleaf pine is better able to sustain dominance in landscapes characterized by frequent fire and the risk of southern pine epidemics than loblolly and shortleaf pines.

The extent of longleaf pine communities is currently reduced to less than 3% of the area occupied when Europeans reached the region (Frost 1993). In addition, the majority of the remaining stands are severely degraded, primarily due to changes in the fire regime. Wildfires currently have little ecological impact on the West Gulf Coastal Plain, and the presence of fire-maintained communities is dependent on prescribed fire.

Typically, fire-maintained longleaf pine communities of the West Gulf Coastal Plain are characterized by a well- developed herbaceous understory of grasses, primarily bluestem species, and a diverse array of forbs (Bridges and Orzell 1989). Most of the plant diversity resides in the very large number of herbaceous species. Reduction in the frequency or effectiveness of fire can rapidly result in invasion of woody species that progressively out-compete the herbaceous flora (Streng et al. 1993). Consequently, many herbaceous species on the West Gulf Coastal Plain are of conservation concern due to past changes in the fire regime. Lack of fire also impacts many included communities, such as pitcher plant bogs, that also support many rare and localized species.

Numerous animal species are also characteristic of longleaf pine communities and are detrimentally impacted by changes in the fire regime (Engstrom 1993, Guyer and Bailey 1993). Northern Bobwhite have declined precipitously in the region due to lack of fire to maintain appropriate habitat. Huntable populations are currently rare. In addition, Henslow’s Sparrow, Bachman’s Sparrow, Louisiana pine snake, and others are declining due to habitat loss and changes in the fire regime.

Most notable among the species declining due to changes in longleaf pine ecosystems is the Red-cockaded Woodpecker (Conner and Rudolph 1989, James 1995). The Red-cockaded Woodpecker is unique among woodpeckers in that it excavates cavities exclusively in living pines. This behavior presents problems during excavation, but provides an opportunity as well. The continued excavation of resin wells in the vicinity of the cavity entrance produces a resin barrier on the bole of the pine that greatly reduces the ability of rat snakes, a potential predator, to reach the cavity (Rudolph and Conner 1990).

From a management perspective, longleaf pine has a number of advantages that make it attractive to forest managers (Boyer 1993, Farrar 1993, Rudolph and Conner 1996). Resistance to fire and southern pine beetles serve to protect investments. In addition, longleaf pine is amenable to a wide variety of silvicultural methods from even-age to uneven-age (especially group selection) that fit a wide variety of management goals.

11 Recent advances in establishing state wide safe harbor agreements for Red-cockaded Woodpeckers are also reducing concerns of managers.

Restoration of longleaf pine ecosystems presents many challenges due to the current limited extent of longleaf pine dominated forests and the poor condition of the associated ecosystem in most of the remaining stands. Restoration opportunities range from replacing longleaf pine on sites where it no longer occurs to restoring the presence and function of the associated ecosystem. The primary method to achieve this latter goal is return of a frequent fire regime. However, opportunities to achieve additional restoration objectives could range from reintroduction of Red-cockaded Woodpeckers to a wide variety of plant species. The important point is that a wide range of restoration opportunities are possible, that are compatible with a variety of management goals.

Speakers in the Ecology Concurrent Session will elaborate on these topics in the afternoon session.

LITERATURE CITED Bridges, E. L. and S. L. Orzell. 1989. Longleaf pine communities of the West Gulf Coastal Plain. Natural Areas Journal 9:246-253. Boyer, W. D. 1993. Regenerating longleaf pine with natural seeding. Proc. Tall Timbers Fire Ecology Conf. 18:299-309. Conner, R. N. and D. C. Rudolph. 1989. Red-cockaded Woodpecker colony status and trends on the Angelina, Davy Crockett and Sabine National Forests. U. S. Dep. Agric., For. Serv. Research Paper SO- 250. Conner, R. N. and D. C. Rudolph. 1995. Losses of Red-cockaded Woodpecker cavity trees to southern pine beetles. Wilson Bull. 107:81-92. Engstrom, R. T. 1993. Characteristic mammals and birds of longleaf pine forests. Proc. Tall Timbers Fire Ecology Conf. 18:127-137. Farrar, R. M. 1993. Growth and yield in naturally regenerated longleaf pine stands. Proc. Tall Timbers Fire Ecology Conf. 18:311-335. Frost, C. C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. Proc. Tall Timbers Fire Ecology Conf. 18:17-43. Guyer, C. and M. A. Bailey. 1993. Amphibians and reptiles of longleaf pine communities. Proc. Tall Timbers Fire Ecology Conf. 18:139-157. Hodges, J. D., W. W. Elam, W. F. Watson, and T. E. Nebeker. 1979. Oleoresin characteristics and susceptibility of four southern pines to southern pine beetle (Coleoptera: Scolytidae) attacks. Can. Entomol. 111:889-896. James, F. C. 1995. The status of the Red-cockaded Woodpecker in 1990 and the prospect for recovery. Pp. 439-451 in D. L. Kulhavy, R. G. Hooper, and R. Costa (eds.). Red-cockaded Woodpecker: Recovery, ecology and Management. College of Forestry, Stephen F. Austin State Univ., Nacogdoches, Texas. Maxwell, R. S. and R. D. Baker. 1983. Sawdust empire, the Texas logging industry, 1830-1940. Texas A&M Univ. Press, College Station, Texas. 228 pp. Platt, W. D., G. W. Evans, and S. L. Rathbun. 1988. The population dynamics of a long lived conifer (Pinus palustris). Amer. Nat. 131:491-525. Rudolph, D. C., H. Kyle, and R. N. Conner. 1990. Red-cockaded Woodpeckers vs rat snakes: the effectiveness of the resin barrier. Wilson Bull. 102:14-22. Rudolph, D. C. and R. N. Conner. 1996. Red-cockaded Woodpeckers and silvicultural practice: is uneven- aged silviculture preferable to even-aged? Wildl. Soc. Bull. 24:330-333. Streng, D.R., J. S. Glitzenstein, and W. J. Platt. 1993. Evaluating effects of season of burn in longleaf pine forests. Proc. Tall Timbers Fire Ecology Conf. 18:227-263. Wahlenberg, W. G. 1946. Longleaf pine: its use, ecology, regeneration, protection, growth and management. School of Forestry, Duke Univ., Durham, North Carolina. Webb, T., III. 1986. Vegetational change in eastern North America from 18,00 to 500 Yr B. P. Pp. 63-69 in C. Rozenweig and R. Dickinson (eds). Climate-vegetation interactions. Office Interdisciplinary Earth Studies and Univ. Corp. Atmospheric Research, Boulder, Colorado.

12 FIRE HISTORY OF THE WEST GULF COASTAL PLAIN

David Jurney (US Forest Service) Rob Evans (The Nature Conservancy) John Ippolito (US Forest Service) Velicia Bergstrom (US Forest Service)

This paper summarizes the known fire history data for the West Gulf Coastal Plain generally, and the National Forests and Grasslands in Texas, specifically. The National Forests are located in eastern Texas, and the Grasslands are located in north-central Texas. They cover approximately 1054.8 square miles (675,100 acres), comprising ca. 2.5% of the forested portion of Texas. Prior to historic land use modifications the forested portion of Texas covered approximately 42,000 square miles (26,880,000 acres). The forests generally consisted of 12% longleaf pine-dominated forest, 71% shortleaf pine-dominated forest, and 17% loblolly-dominated forest (Swanson 1995, Pool et al. 1975).

The East Texas Pineywoods grade from almost pure stands of pines, to mixed pine-hardwoods, to almost pure hardwoods; as shown by the Texas General Land Office surveys. The frequency of pine trees increases from the Trinity River to the Sabine River. Pines comprise over half of the tree species in upland and slope settings. Pines are found in floodplains; but seldom reach dominant status there. Small and large prairies are found in uplands, on slopes, and in floodplains; extending to the easternmost parts of Texas. Forests in the natural, “old growth” state consisted of multiaged pines, with various amounts of hardwoods, and fire tolerant grass understory (Landers and Boyer 1999). Fire intolerant species, such as bois d’arc, were confined to wet floodplain prairies where fires were infrequent (Jurney 1994, Weniger 1996).

By combining the information derived from the Weakly Bog and Boriack Bog on the Blackland Prairie-Post Oak Savannah border in east-central Texas, a 16,000-year pollen sequence has been developed (Bousman 1991), that identifies changes in the proportion of 23 tree taxa and grass pollen over the period of record. These data indicate that nine major plant communities have been present in the West Gulf Coastal Plain, ranging in percentage of canopy cover from 10-47%. Of particular noteworthiness is the presence of spruce (Picea glauca) in the earliest portions of this record. Spruce declined from 6.5% of pollen 16,000 YBP, to 0.5% around 10,000 YBP; after which it vanished. Oaks were present throughout the pollen sequence. Pines were highest during the Pleistocene, but reached the current level ca. 9000 YBP.

Fire History Database. The fire records for Texas are unpublished, available only in computerized format from the Texas Forest Service and the U.S. Forest Service. Law enforcement and fire management personnel identify the causes of fires, but time, money, and flames often prevent firm identification. Attribution to lightning is avoided unless obvious, thereby under enumerating the true number of lightning ignitions. As early as 1898, the Forestry Bureau of the U.S. Department of Agriculture began a cooperative program with Texas lumber companies to restore the forests and develop a sustained yield strategy (Baker n.d., Ippolito et al. 1999). From 1898-1915, there were no data recorded on fires. From 1916-1921, fire data were reported for Texas, but were apparently restricted to the 697,000 acres considered for purchase from private timber companies (Baker n.d.). In 1922, the Texas Forest Service joined the U.S. Forest Service in reporting fires from limited portions of Texas. Gary Lacox, of the Texas Forest Service (TFS), reports that surveys of volunteer fire departments indicate that for every fire responded by TFS and the U.S. Forest Service, there are nine fires that go unenumerated. The geographic extent of the USFS/TFS enumeration is estimated to cover less than 4,666 square miles (2.98 million acres) by excluding the volunteer fire department responses over the entire forested area, ca. 42,000 square miles (26,880,000 acres) of Texas.

The fire records indicate a moderately high frequency of lightning-caused fires. These data yield a minimum estimate of 23.72 lightning fires/year in the areas that the Texas Forest Service and the U.S. Forest Service responded. Lightning-caused fires occurred every year, with a minimum estimate of 1,945 lightning fires, and a maximum estimate of 17,505 lightning fires in an 82-year period. Lightning fires peak in frequency from March to October, but have been recorded every month of the year. There are approximately eight lightning fires/million acres/year in the forested portion of Texas

13 Fire Scars in Trees. The oldest known stand of longleaf pines in East Texas is Pine Park, near Hemphill in Sabine County. This relict 4-acre parcel yielded a tree-ring chronology from 1755-1995, indicating that severe droughts return at least every 11.5 years, minor droughts every 2-4 years, and fires at frequencies of 1.5 years. An examination of mixed loblolly and longleaf tree stumps and logs from the 1998 Blowdown on the Sabine N.F. indicated fire scars at 3.75 year return intervals, dating from 1855-1930. Fire scars from longleaf trees adjacent to the Concord Cemetery on the Angelina National Forest indicate a fire return interval of 7 years. Based on fire scars in trees, the natural and cultural fire return interval for East Texas ranges between 1.5 and 7 years; corroborating the frequencies estimated by Frost (1998).

Dendroclimatology. Comparison of tree growth and the Palmer Drought Severity Index (PDSI) (Palmer 1965) has been conducted for the continental United States (Stahle et al. 1984, 1985; Cook et al. 1994, 1997), showing the patterns of drought over the last 300 years. These data are graphically presented at the Internet website www.ngdc.noaa.gov/paleo/pdsiyear.html. The data for Texas (1916-1995) were compared to the U.S. Forest Service period-of-record for lightning fires (1916-1990). Severe lightning fire years correspond to 56% of the years of below normal (drought) PDSI. Independent research by Don Latham has found a low correlation between lightning fires and positive PDSI.

CONCLUSIONS A cursory examination of the fire data indicates that the time between extreme lightning fire years is shortening. Also, there appears to be an increase in the average numbers of lightning fires. The most extreme year, 1956, is closely matched by 1996 and 2000.

Historically, ecologists have assumed that plant and animal communities move through a series of succession stages toward steady state, or climax, equilibrium. Since 1990, ecologists analyzing data from more intensive, long-term studies have shifted from this perspective, requiring a total reconsideration of strategies of conservation and resource management (Stevens 1990, Oliver and Larson 1996:5, Evans 1997). The role of human and natural disturbances is seen to be increasingly important in determining forest structure and species composition (Delcourt et al. 1998, Barden and Woods 1974). Fire is only one of many sources of disturbance. Fires that occur without suppression activities burn tens to hundreds of thousands of acres. Even tropical rain forests are susceptible, particularly during weather events such as drought or El Nino/Southern Oscillation. Evidence of past fires has been found in all forest regions and all forest types around the globe (Stevens 1990:100).

As this study suggests, the combination of drought and lightning fire provided a natural setting for extensive wildfires. By adding the human factor, fires played a much greater role in the fluctuating mosaic of forests in Texas than previously acknowledged. Modern fire suppression, reduced timber management, and the establishment of wilderness areas where no treatments are allowed all exacerbate the fire danger we now find in the forests of the West Gulf Coastal Plain.

REFERENCES CITED Baker, Robert D. n.d. Timbered Again, the History of the National Forests in Texas. Unpublished manuscript on file, Supervisor’s Office, National Forests and Grasslands in Texas, Lufkin. Barden, L.S. and F.W. Woods. Characteristics of Lightning Fires in Southern Appalachian Forests. Tall Timbers Fire Ecology Conference 13:345-361. Bousman, C. Britt. Paleoenvironments. In Excavations at the Bottoms, Rena Branch, and Moccasin Springs Sites, Jewett Mine Project, Freestone and Leon Counties, Texas, edited by Ross C. Fields, pp. 21-36. Prewitt and Associates, Report of Investigations 82, Austin. Cook, E. R., D. M. Meko, D. W. Stahle, and M. K. Cleveland. 1994. Tree-Ring Reconstruction of Past Drought Across the Coterminous United States: Tests of the Regression Method and Calibration/Verification Results. Proceedings of the International Conference on Tree Rings, Environment, and Humanity: Relationships and Processes. Tucson. Cook, Edward R., J.S. Glitzenstein, P.J. Krusic, and P.A. Harcombe. 1997. Phylogenetic Signals in Annual Tree-Ring Chronologies from the Big Thicket Region of East Texas. Ms on file Tree-Ring Laboratory, Lamont-Doherty Earth Observatory, Palisades, NY.

14 Delcourt, P.A., H.R. Delcourt, C.R. Ison, W.E. Sharp, and K.J. Gremillion. 1998. Prehistoric Human Use of Fire, The Eastern Agricultural Complex, and Appalachian Oak-Chestnut Forests: Paleoecology of Cliff Palace Pond, Kentucky. American Antiquity 63(2):263-278. Evans, Robert E. 1997. Distribution and Composition of Historical Forests in two Eastern Texas Counties, With Specific Reference to Longleaf Pine. M.S. thesis, Biology, Stephen F. Austin University, Nacogdoches, Texas. Frost, Cecil C. Presettlement Fire Frequency Regimes of the United States: A First Approximation. In Teresa L. Pruden and Leonard A. Brennan (eds.) Fire in Ecosystem Management: Shifting the Paradigm from Suppression to Prescription, pp. 70-81. Tall Timbers Fire Ecology Conference Proceedings, No. 20. Tall Timbers Research Station, Tallahassee, FL. Ippolito, John, V. Bergstrom, D. Jurney, and E. Sherman. 1999. An Historic Context for the Early Logging Industry in East Texas: A Planning Document for the National Forests and Grasslands in Texas. Ms on file U.S.D.A. Forest Service, Lufkin, TX. Jurney, David H. 1994. The Original Distribution of Bois D’arc. Part I: Texas. Caddoan Archeology Newsletter 5(2):6-13. Landers, J. Larry and William D. Boyer. 1999. An Old-Growth Definition for Upland Longleaf and South Florida Slash Pine Forests, Woodlands, and Savannas. General Technical Report SRS-29. U.S. Department of Agriculture, Forest Service, Southern Research Station. Oliver, Chadwick D. and B.C. Larson. 1996. Forest Stand Dynamics. John Wiley and Sons, New York. Palmer, W.C. 1965. Meteorological Drought. Weather Bureau Research Paper 45. U.S. Department of Commerce, Washington D.C. Pool, William C., Edward Triggs, and Lance Wren. A Historical Atlas of Texas. The Encino Press, Austin. Stahle, David W., E.R. Cook, and J.W.C. White. 1985. Tree-Ring Dating of Baldcypress and the Potential for Millenia-Long Chronologies in the Southeast. American Antiquity 50(4):796-802. Stahle, David W., J.G. Hehr, G.G. Hawks, M.K. Cleaveland, and J.R. Baldwin. 1984. Tree-Ring Chronologies for the Southcentral United States. Report Submitted to the National Science Foundation. Grant ATM-8006964. Stevens, W.K. 1990. New Eye on Nature: The Real Constant is Eternal Turmoil. New York Times. Science Article. Tuesday, July 31, 1990. B5-B6. Swanson, Eric R. Geo-Texas: A Guide to the Earth Sciences. Texas A&M University Press, College Station. Weniger, Del. 1996. Catalpa (Catalpa Bignoides, Bignoniaceae) and Bois D’Arc (Maclura Pomifera, Moraceae) in Early Texas Records. Sida 17(1):239-242.

15 LONGLEAF PINE GROWTH AND YIELD REFERENCES PLUS SELECTED OTHERS

Ralph S. Meldahl (School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849) John S. Kush (School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849)

Across the historical range of longleaf pine, less than 10% of the undeveloped lands previously occupied by longleaf ecosystems are currently in public ownership. The remainder is owned by private entities ranging from forest industries to private, non-industrial landowners. Any significant recovery of longleaf is dependent on the participation of the private sector. Although experience has indicated that economic return is often not a primary motivator for all of these landowners, it is usually a factor in investment decisions. Certainly, for forest industry, and many other investors, the prospect of positive returns on investment is extremely important. Motivation for private non-industrial landowners is as varied as they are, but it is unusual for even that group to not consider profitability of their forest investment, even if it is for a future generation.

The resurgence of interest in re-establishment of longleaf pine forests and restoration of longleaf ecosystems has outpaced the growth in our knowledge of some important aspects of that effort. One major area requiring more knowledge is the need for models to reliably project growth and, ultimately, economic value of the longleaf itself. There is limited data available for natural stands of longleaf which may be extrapolated to yield some estimate of potential growth of planted stands, but there is a great deal of uncertainty when gains in seedling quality, competition control, fertilization, and silvicultural techniques are factored in. Much of the re-establishment of longleaf taking place today is occurring on old fields and pastures. At least half of that planting is being done with containerized seedlings, usually employing both site-preparation and follow- up competition control to improve survival and accelerate growth. Credible economic analysis of these investments is impossible without some reliable estimate of expected growth.

Using the best of today’s technology and science, many foresters feel that longleaf can grow competitively with or even exceed the growth of other southern pine species on many sites. If that is true, and if markets continue to award quality wood products, particularly utility poles, with premium prices, longleaf can be an extremely profitable venture. Private industrial and non-industrial landowners should respond positively to that possibility and make longleaf a vital part of their portfolio.

The following is a list of references related to longleaf pine growth and yield. With the renewed interest and increased planting of longleaf pine, it is hoped that data will be collected to improve the models available.

REFERENCES CITED Compiled by Ralph Meldahl and John Kush (School of Forestry and Wildlife Sciences, Auburn University) for The Third Longleaf Alliance Conference, “Forestry for Our Future”, Alexandria, LA, October 16-18, 2000 Baldwin, V.C., Jr., and Saucier, J.R. 1981. Weight and volume prediction of unthinned, planted longleaf pine.P. 27-38 in Proc. 1981 South. For. Biomass Workshop. Belle W. Baruch For. Sci. Inst., Clemson Univ. Baldwin, V.C., Jr., and Polmer, B.H. 1981. Taper functions for unthinned longleaf pine plantations on cutover West Gulf sites. P. 156-163 in Proc. First Bien. South. Silvicultural Res. Conf. USDA For. Serv. Gen. Tech. Rep. SO-34. Baldwin, V.C., Jr., and Saucier, J.R. 1983. Aboveground weight and volume of unthinned, planted longleaf pine on west Gulf forest sites. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-191, 25 p. Baldwin V.C., Jr., Leduc, D.J., Peterson, K.D., and Parresol, B.R. 1998. Basic growth relationships in thinned and unthinned longleaf pine plantations. In: Proceedings of the 2nd Longleaf Alliance Conference; 1998 November 17-19: Charleston, SC. Longleaf Alliance Report No. 4. Auburn University, AL; Longleaf Alliance: 49-51 Bethune, J.E., and Roth, E.R. 1960. Source of seed affects growth of longleaf pine. USDA For. Serv., Southeast. For. Exp. Stn. Res. Note 146, 2 p. Bower, D., and Schmitt, D. 1970. Volume tables for young loblolly, slash, and longleaf pines in plantations in south Mississippi. USDA For. Serv. Res. Note SO-102, 6 p.

16 Boyer, W.D. 1974. Longleaf pine seedling mortality related to year of overstory removal. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-181, 3 p. Boyer, W.D. 1974. Impact of prescribed fires on mortality of released and unreleased longleaf pine seedlings. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-182, 6 p. Boyer, W.D. 1968. Foliage weight and stem growth in longleaf pine. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-86, 2 p. Boyer, W.D. 1979. Mortality among seed trees in longleaf pine shelterwood stands. South. J. Appl. For. 3(4):165-167. Boyer, W.D. 1981. Site and stand factors affecting height growth curves of longleaf pine plantations. In Proc. First Biennial South Silv. Res. Conf., Atlanta, GA Nov. 6-7, 1980, J. P. Barnett, ed. p.184-187, USDA Forest Serv. Gen. Tech. Rep. SO-34. Boyer, W.D. 1983. Growth of young longleaf pine as affected by biennial burns plus chemical or mechanical treatments for competition control. P. 62-65 in Southeast. For. Exp. Stn., USDA For. Serv. Gen. Tech. Rep. SE-24. Boyer, W.D. 1983. Variations in height-over-age curves for young longleaf pine plantations. For. Sci. 29(1):15-27. Boyer, W.D. 1987. Volume growth loss: a hidden cost of periodic prescribed burning in longleaf pine? South. J. Appl. For. 11(3):154-157. Boyer, W.D. 1988. Effects of site preparation and release on the survival and growth of planted bare-root and container-grown longleaf pine. Georgia forest research paper. Aug 1988. (76) 9 p. ill. [S.l.] : Research Division, Georgia Forestry Commission. Boyer, W.D. 1992. Season of burn and hardwood development in young longleaf pine stands. In: Proceedings of the 7th biennial southern silvicultural research conference; 1992 November 17-19; Mobile, AL. Gen. Tech. Rep. SO-93 New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station; 511-515. Boyer, W.D. 1993. Long-term development of regeneration under longleaf pine seedtree and shelterwood stands. South. J. Appl. For. 17(1):10-15. Boyer, W.D. 1994. Eighteen years of seasonal burning in longleaf pine: effects on overstory growth. In: Proc. 12th International Conference on Fire and Forest Meteorology, 1993 Oct. 26-28, Jekyll Island, GA. Society of American Foresters, p. 602-610. Bethesda, MD. 796 p. Boyer, W.D. 1996. Longleaf pine can catch up. In: Kush, John S., comp. Proceedings of the First Longleaf Alliance conference – Longleaf Pine: a regional perspectie of challenges and opportunities. 1996. September 17-19; Mobile, AL: [Auburn University, AL]. Longleaf Alliance Report No. 1. 28-29. Boyer, W.D. 2000. Long-term effects of biennial prescribed fires on the growth of longleaf pine. P. 18-21 in W. Keith Moser and Cynthia F. Moser (eds.). Fire and forest ecology: innovative silviculture and vegetation management. Tall Timbers Fire Ecology Conference Proceedings. No. 21. Tall Timbers Research Station, Tallahassee, FL. Cao, Q.V., and Pepper, W.D. 1986. Predicting inside bark diameter for shortleaf, loblolly, and longleaf pines. South. J. Appl. For. 10(4):220-224. Cao, Q.V., Baldwin, V.C. Jr., and Lohrey, R.E. 1997. Site index curves for direct-seeded loblolly and longleaf pines in Louisiana. South. J. Appl. For. 21(3):134-138. Curlin, J.W. and Box, B.H. 1961. Estimating past annual height growth of slash and longleaf pines from length of internodes. J. For. 59:372-373. Derr, H.J. and Enghardt, H.G. 1969. Growth in a young managed longleaf pine plantation. J. For. 67(7):501-504. Farrar, R.M., Jr. 1974. Growth and yield of young longleaf pine. P. 276-287 in Proc. Symp. Manage. Young Pines. USDA For. Serv., Southeast. Area, State & Priv. For. Farrar, R.M., Jr. 1978. Silvicultural implications of the growth response of naturally regenerated even-aged stands of longleaf pine (Pinus palustris Mill.) to varying stand age, site quality, and density and certain stand structure measures. Ph.D. Thesis, Univ. Ga., Athens. 132 p. Farrar, R.M., Jr. 1979. Volume production of thinned natural longleaf. P. 30-48 in W.E. Balmer Ed. Proc. Longleaf Pine Workshop, Mobile, Al. Oct. 17-19, 1978, USDA For. Serv. Tech. Pub. SA-TP3, SE Area, State & Priv. For., Atlanta, Ga. Farrar, R.M., Jr. 1979. Growth and yield predictions for thinned stands of even-aged natural longleaf pine. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-156, 78 p.

17 Farrar, R.M., Jr. 1981. A site-index function for naturally regenerated longleaf pine in the East Gulf area. South. J. Appl. For. 5(3):150-153. Farrar, R.M., Jr. 1981. Cubic-foot volume, surface area, and merchantable height function for longleaf pine trees. USDA For. Serv. Res. Pap. SO-166. Farrar, R.M., Jr. 1985. Crown ratio used as a surrogate for form in a volume equation for natural longleaf pine stems. In: Shoulders, E., ed. Proceedings - third biennial southern silvicultural research conference; 1984 November 7-8; Atlanta, GA: Gen. Tech. Rep. SO-54. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station:429-435. Farrar, R.M., Jr. 1985. Predicting stand and stock tables from a spacing study in naturally regenerated longleaf pine. Res. Pap. SO-219. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station. 28 p. Farrar, R.M., Jr. 1985. Volume and growth predictions for thinned even-aged natural longleaf pine stands in the east Gulf area. Res. Pap. SO-220. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station. 171 p. Farrar, R.M., Jr. 1987. Stem-profile functions for predicting multiple-product volumes in natural longleaf pines. South. J. Appl. For. 11(3):161-167. Farrar, R.M., Jr. 1990. Predictions of volume and volume growth in naturally-regenerated longleaf pine stands. In: Proc. Symposium on the Management of Longleaf Pine. 1989 April 4-6, Long Beach, MS. USDA Forest Service, General Technical Report SO-75, p. 170-192. Southern Forest Experiment Station, New Orleans, LA. 293 p. Farrar, R.M., Jr., and Boyer, W.D. 1991. Managing longleaf pine under the selection system--promises and problems. In: Coleman, S.S.; Neary, D.G. eds. Proceedings of the sixth biennial southern silvicultural research conference; 1990 October 30-November 1; Memphis, TN: Gen. Tech. Rep. SE- 70; Asheville, NC: U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station:357-368. Farrar, R.M., and Matney, T.G. 1994. A dual growth simulator for natural even-aged stands of longleaf pine in the South's east Gulf region. South. J. Appl. For. 18(4):147-155. Gilliam, F.S. 1991. Stand structure and longleaf pine regeneration of an old-growth longleaf pine forest under chronic fire exclusion. Bull. Ecol. Soc. Amer. Abstract 72(2):124. Goodwin, O.C. 1980. Fertilization increases straw production and tree growth of longleaf pine stand in North Carolina. Raleigh: Division of Forest Resources, 1980. 6 p.; 28 cm. (Forestry Note no. 43, Forest Service, NC, Division of Forest Resource). Goodwin, O.C. 1982. Fertilization of longleaf pine stand increases straw production and tree growth yields up to three years. Forestry Note No.56, Div. of Forest Resource, Raleigh, NC. 5p. Harrington, C.A. 1990. PPSITE-A new method of site evaluation for longleaf pine: Model development and user’s guide. USDA For. Serv. Res. Gen. Tech. Rep. SO-80. Haywood, J.D. and Grelen, H.E. 2000. Twenty years of prescribed burning influence the development of direct-seeded longleaf pine on a wet pine site in Louisiana. South. J. Appl. For. 24(2):86-92. Jackson, L.W.R. 1968. Growth and specific gravity of unthinned longleaf, slash pine (Pinus palustris, P. elliottii) in Georgia Piedmont. Nav. Stores Rev. Terpene Chem. 78(6): 10-11. Sept. 1968. 99.81 N22. Kush, J.S., Boyer, W.D. Meldahl, R.S., and Ward, G.A. 1998. Precommercial thinning intensity in longleaf pine: effect on product volume and value. In: Proceedings of the 2nd Longleaf Alliance Conference; 1998 November 17-19; Charleston, SC. Longleaf Alliance Report No. 4. Auturn University, AL: Longleaf Alliance: 106-108 Lauer, D.K. 1987. Seedling size influences early growth of longleaf pine. Tree Planters' Notes, 38(3):16- 17. Lohrey, R.E. 1974. Growth and yield in a longleaf pine plantation. P. 86-88 in Proc. Symp. Manage. Young Pines. USDA For. Serv. Southeast Area, State & Priv. For. Lohrey, R.E. 1974. Growth of longleaf pine plantations after initial thinning. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-175, 5 p. Lohrey, R.E., and Bailey, R.L. 1977. Yield tables and stand structure for unthinned longleaf pine plantations in Louisiana and Texas. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-133, 53 p. Lohrey, R.E. 1979. Predicted growth of longleaf pine planted on cutover forest sites in the West Gulf. P. 54-63 in W.E. Balmer Ed. Proc. Longleaf Pine Workshop, Mobile, Al. Oct. 17-19, 1978, USDA For. Serv. Tech. Pub. SA-TP3, SE Area, State & Priv. For., Atlanta, Ga. Lohrey, R.E. 1983. Stem volume predictions and crown characteristics of thinned longleaf pine plantations. P. 338-343 in Southeast. For. Exp. Stn., USDA For. Serv. Gen. Tech. Rep. SE-24.

18 Maple, W.R. 1976. How to estimate longleaf seedling mortality before control burns. J. For. 74(8):517-518. Murphy, P.A., and Farrar, R.M. Jr. 1989. A new mortality (or survival) function for longleaf pine plantations. USDA For. Gen. Tech. Rep. SO-74. U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station. 1989. (74) p. 427-432. Page, H.H., Jr. 1990 Six-year growth response of longleaf pine (Pinus palustris Mill.) seedlings to varying intensities of site preparation treatments. Forestry Research Report-Smurfit Group, plc 1990. (No. 4): 10 pp. Quicke, H.E., Meldahl, R.S., and Kush, J.S. 1994. Basal area growth of individual trees: a model derived from a regional longleaf pine growth study. For. Sci. 40(3):528-542. Saucier, J.R., Phillips, D.R., and Williams, J.G., Jr. 1981. Green weight, volume, board-foot, and cord tables for the major southern pine species. Ga. For. Res. Counc. Res. Pap. 19, 63 p. Schmitt, D. and Bower, D. 1970. Volume tables for young loblolly, slash and longleaf pines in plantations in south Mississippi. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-102, 6 p. Schroeder, J. G., Taras, M.A., and Clark, A., III. 1975. Stem and primary product weights for longleaf pine sawtimber trees. Southeast. For. Exp. Stn., USDA For. Serv. Res. Pap. SE-139, 15 p. Shain, W.A., and Jenkins, W.B. 1974. Growth and yield of loblolly, longleaf, and slash pines in the South Carolina sandhills. Clemson Univ. Dep. For. Bull. 15, 3 p. Shain, W.A., and Helms, J.R. 1980. Growth and yield of loblolly, slash, and longleaf pines in the South Carolina Sandhills: third measurement. (Final report) Department of Energy, Washington, DC. Report no. DOE/SR/10706-1. Shoulders, E. 1967. Growth of slash and longleaf pines after cultivation, fertilization, and thinning. South. For. Exp. Stn., USDA For. Serv. Res. Pap. SO-59, 3 p. Smith, L.F. 1961. Growth of longleaf pine seedlings under large pines and oaks in Mississippi. USDA For. Serv., South. For. Exp. Stn., Occas. Pap. 189, 4 p. Smith, L.F., and Smith, H.D. 1963. Growth of slash, loblolly, and longleaf pines on cultivated sites. Tree Planters' Notes 59:1-2. Somers, G.L., and Farrar, R.M., Jr. 1991. Biomathematical growth equations for natural longleaf pine stands. For. Sci. 37(1):227-244. Shaw, D.J., Meldahl, R.S., Kush, J.S., Quicke, H.E., and Farrar, R.M., Jr. 1991. Pole availability from naturally regenerated longleaf pine stands: preliminary data. In: Coleman, S.S.; Neary, D.G. eds. Proceedings of the sixth biennial southern silvicultural research conference; 1990 October 30-November 1; Memphis, TN: Gen. Tech. Rep. SE-70; Asheville, NC: U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station:260-264. Sparks, R.C., Linnartz, N.E. and Harris, H.E. 1980. Long-term effects of early pruning and thinning treatments on growth of natural longleaf pine. Southern Journal of Applied Forestry. 4:77-79 Thomas, C.E., Parresol, B.R., Le, K.H.N., and Lohrey, R.E. 1995. Biomass and taper for trees in thinned and unthinned longleaf pine plantations. South. J. Appl. For. 19(1):29-35. van Lear, D.H., Saucier, J.R., and Williams, J.G., Jr. 1977. Growth and wood properties of longleaf pine following silvicultural treatments. Soil Sci. Soc. Am. J. 41(5):989-992. West, D.C., Doyle, T.W., Tharp, M.L., Beauchamp, J.J., Platt, W.J., and Downing, D.J. 1993. Recent growth increases in old-growth longleaf pine. Can. J. Forest Res. 23(5):846-853. White, F.M., Kunselman, M.B., Robbins, D.H., Crain, W.T., Hood, C.A., Precythe, H.W. and Moehler, F.A. 1988. Establishment and growth of longleaf pine on droughty sites in North Carolina. North Carolina Forest Service, Forestry Note No. 61. Division of Forest Resources, Raleigh, NC 25+10 p.

19 SILVOPASTURE - AN AGROFORESTRY PRACTICE FOR LONGLEAF PINE?

James L. Robinson (USDA, NRCS, National Agroforestry Center)

The traditional model of agriculture places cropland, forest land and pastureland in separate fields on the farm. They are separated either by fence or management. With such a model the increase recognition and establishment of longleaf pine must rely on the conversion of pasture or cropland to forest land or the reforestation of cut over areas following final harvest to a longleaf pine forest.

There is another option that warrants investigation. The integration of longleaf pine into existing cropland or pasture through an agroforestry management system. Agroforestry is the integration of forestry technologies and agricultural technologies for the improvement of economic or environmental values. Agroforestry creates a productive, profitable, healthy, sustainable commercial land-use production system.

This paper will emphasize the agroforestry system known as silvopasture or the integration of livestock, forage and forest production. It is not just the traditional opportunistic woodland grazing that gleans forage production after a thinning until the stands became too dense and forage too scarce to be of an economic value. These systems often had periods of overgrazing where environmental impacts could have negative impacts on wildlife, trees, and livestock production. Silvopasture is intentional, integrated, interactive, and intensive. Through careful management of stand and canopy density, silvopasture can optimize timber and forage growth and provide easy equipment access for the management of both the trees and forage.

Silvopasture systems can be developed from existing, natural or plantation, pine stands or from pasture field. For example, in much of its range, longleaf stands often had an understory or native forbs and grasses. Through natural succession and low level forest management hardwood encroachment often occurred. Removal of the hardwood timber and management of the pine and forage can utilize grazing management to prevent hardwood encroachment and maintain the pine savanna. This may require prescribed burning to clean up the site, seed bed preparation and the reestablishment of desired forage species. Stand basal area would typically be 20 to 30 sq. feet less than that recommended for plantation pine management depending on the grass species being utilized in the understory. The advantage of starting with an established pine plantation or mixed pine hardwood stand is that the silvopasture system can be implemented as soon as the hardwoods are moved, pines are thinned, and the grass is established.

Silvopasture systems developed from existing pastures take a few years before the system is ready for grazing. It is not recommended that this area be grazed until the terminal bud of the tree is above the reach of the livestock, usually 2-3 years. Forage can be harvested at this time for hay. The ease of haying is dependent upon the tree planting configuration. The pines are established in wide spaced rows typically at a planting density of 300 to 400 trees per acre. There are any number of planting arrangements that can be used to achieve this planting density. Trees spaced evenly across a field at a 10 ft. by 12 ft. spacing will give you a planting density of 363 trees per acre. A set of two closely spaced rows of trees with a spacing at 6 ft. by 10 ft. and then a wide alley of 30ft followed by another set of rows will also provide a planting density of about 363 trees per acre.

Management is the key to successful silvopasture implementation, intensive management. Trees generally have little impact on forage production until shading becomes dense enough to limit sunlight to the understory. Thinning of trees is scheduled to reduce canopy shade and tree competition when understory forage production begins to decline. When the trees' combined canopy exceeds 35 to 45%, forage production of warm-season grasses begins to decline. However, there are significant differences among warm-season grasses. For instance, Pensacola bahiagrass and Coastal bermundagrass were shown to produce more under tree canopy cover than dallisgrass and carpetgrass. Continuous observation is important in making adjustments in the management strategy. For cool season grasses, shade tolerance of some species may exceed 60% canopy and still produce good forage yields. Depending upon the species of grass, tree thinning needs to be conducted to keep canopy cover below the maximum shade tolerance. With proper establishment densities, the first thinning should be planned around 10 to 15 years of age for pulp or small round wood. Successive thinning can be scheduled about every five years until final harvest at approximately 30 to 45

20 years. This schedule will vary some depending upon the productivity of the site, the species of trees, and the targeted final wood products.

Widely spaced trees delay canopy closure benefiting forage crops but the "open grown" trees develop large lateral branches that will reduce wood quality if the trees are not pruned. Pruning is essential. The object of pruning is to confine the knots created by these lateral branches to a small diameter (four inches) core wood thereby producing high quality, knot free wood on the outer diameter of the tree stem. Pruning should be initiated when the crop trees reach 15 to 20 feet and/or the stump diameter reaches five inches at a height six inches above the ground. Pruning should strive to remove all of the branches where the trunk diameter is greater than four inches, but never remove more than 50% of the live canopy. Periodically schedule pruning operation until the tree bole is pruned up to 18 feet. Each successive pruning operation proceeds up the main stem to a four-inch diameter core but removes no more than 1/2 of the total crown while maintaining a live crown equal to 1/3 of the tree height. Follow the local guidelines developed by the state forestry agency, NRCS, or extension service for proper pruning techniques.

Grazing management is extremely important to prevent damaging impacts to trees. Correlation of increased diseases with livestock grazing in pine has generally been associated with root damage due to animal concentration areas and overgrazing. Continuous or full-season grazing is not recommended for silvopasture systems. A planned grazing system in which multi-grazing units are rested and grazed in a planned sequence is an essential part of a silvopasture system. The grazing management plan should provide for needed water and maintain an adequate balance between livestock numbers and forage production. As with any managed grazing system, soil amendments should be applied as needed to maintain desired forage production levels. Work done at Louisiana State University's Hill Farm Research Station has also shown an increased response in tree growth. Good herd and livestock management is essential. Livestock do not typically browse Southern pines when adequate forage quality is available. Livestock that are not getting adequate nutrition and minerals will try to meet their needs from whatever source possible and the incidence of browse damage to trees is then likely.

Why would a forest landowner or livestock farmer be interested in silvopasture? Livestock price reports over the last 10 years show very few years where farmers received profitable prices for their livestock. Livestock do, however, provide an improved annual cash flow over the typical forest stand with periodic harvests. Conversely, the increase in timber value has been rather dramatic over the last ten years. Close monitoring of forage, livestock, and timber performance will provide economic benefits. Research at Louisiana State University's, Hill Farm Research Station has shown that there was no statistical difference when comparing forage production and animal performance between open pastures and silvopastures. There was also very little difference in biomass production between the pine plantation and the silvopasture system although the open grown "silvopasture" pine had a greater percentage of the biomass removed as sawtimber. Clason's work at the Hill Farm Research Station showed a marked advantage for the silvopasture management alternative economically. The pine plantation provided an internal rate of return of 8.5%, the pasture 6.4% and the silvopasture 12.1%.

Currently, only anecdotal references address the increased benefits to wildlife but the diversity of plant species that exist in silvopasture compared to that in pasture or dense pine stands suggest the potential for enhancing wildlife habitat for some species such as deer and turkey warrants consideration. And while only an observation, most people think silvopasture systems produces a very aesthetically pleasing landscape.

CONCLUSION While most of the existing work with silvopasture has been done with loblolly and slash pine, longleaf growth, quality and habitat makes it a desirable species to explore for silvopasture implementation. Silvopasture can: Increase potential acreage for longleaf reestablishment by planting in pastures that will remain for forage production. Increase the product diversity removed from an acre Provide improved cash flow Provide improved economic returns

21 Potentially provide opportunities for improved wildlife habitat management And create an aesthetically pleasing landscape.

REFERENCES "From a Pasture to A Silvopasture System." Robinson, J.L. and T. R. Clason. 2000. USDA NAC Agroforestry Note 22. Pp. 1-4. "From a Pine Forest to a Silvopasture System." Clason, T.R. and J.L. Robinson. 2000. USDA NAC Agroforestry Note 18. Pp. 1-4. " An Ongoing Study to Understand Tree, Forage, and Livestock Systems." Clason, T.R. 1998. Inside Agroforestry 12(2): 1,5. "Agroforestry, Proceedings Southeastern regional Conference Grazing Lands and People," Pearson, Henry A. 1984. Editors Merkle, Dan; Carter, Roy; Artz, John l.; December 10-12; Atlanta, Ga. 72-79. "Development of Silvopasture Systems in the Northern Temperate Zone." Clason, T. R. 1996 Inside agroforestry 10(2): 3-7. "Double vs. Single-Row Pine Plantations for Wood and Forage Production." Lewis, Clifford E., Etal. 1985 Southern Journal of Applied Forestry, Vol. 9, No. 1. 55-60. "Economic Implication of Silvipastures on Southern Pine Plantations." Clason, R.R. 1995. Louisiana Agricultural Experiment Station, Agroforestry Systems 29:227-238. "Integration of Pines, Pastures and Cattle in South Georgia, USA," Lewis, Clifford E., Etal. 1983. Agroforestry Systems, 1: 277-297. "Managing Pine Trees and Bahiagrass for Timber and Cattle Production." Byrd, Nathan A., Lewis, Clifford 1983. USDA Forest Service, General Report R8-GR2. Pp. 1-9. "Silvopastoral Practices Sustain Timber and Forage Production in commercial Loblolly Pine Plantations of Northwest Louisiana USA." Clason, T. R. 1999. Agroforestry Systems 44:293-303. "Timber-Pasture Management Enhances Productivity of Loblolly Pine Plantations." Clason, T.R. 1996. Louisiana Agriculture 39(2): 14-16.

22 DOES LONGLEAF MAKE DOLLARS AND SENSE?

Rhett Johnson, (Solon Dixon Forestry Education Center, Andalusia, AL) Bob Franklin (Forestry & Wildlife, Clemson University Cooperative Extension Service)

With interest in longleaf pine at its highest point in decades, maybe ever, land owners and managers are asking what kind of investment it actually is. The answer surprises some, but there is every reason to expect very positive returns on investment and in a reasonably short time span. Lumbermen have long realized the value of longleaf products like high quality straight-grained dimensional lumber and strong durable poles. The market continues to recognize this quality by paying top prices for these products. For years, however, longleaf was regarded as a poor investment for a couple of reasons. First, it was considered a difficult species to plant. If it was established successfully, a lengthy period in the grass stage before it initiated height growth extended the period before income could be earned, gaining longleaf a reputation for slow growth. The tree was also often relegated to Alongleaf@ sites, usually deep dry sands where growth was indeed slow, as it would have been for any species. Recent developments in nursery techniques, management practices, and markets have made that prognosis dated. Better quality bare root seedlings and containerized seedlings have taken much of the risk out of planting longleaf. We have learned much about handling and planting longleaf seedlings in the past several years as well. These gains, coupled with increased knowledge about the role of competing vegetation and the development of selective herbicides to control it, have made it possible to shorten and in some cases eliminate the grass stage. That accomplished, we have learned that longleaf is not, as often reported, a slow grower, only a slow starter. Research has shown that, once established on average and poor sites, it will catch and pass faster starting loblolly or slash pine in a reasonable time, 12-15 years on poor sites and 25-30 years on average sites. On very good, productive sites, it takes longer to catch up, often outside a reasonable investment period if return on investment is the only measure used. One consideration often overlooked is that the growth rate of wood volume is not the only or even the most important measure of the value of a forestry investment. The more important measure is the growth rate in value or dollars. Remember that longleaf products return a premium and value is actually growing at a faster rate than volume.

All investment analysis must be based on assumptions or projections of future performance. Forestry investments are no different. The accuracy of these projections is critical to the accuracy of the analysis. The basic information needs are growth and yield projections. Growth is projected in terms of volume of wood produced and yield in terms of products grown and in what proportions. Unfortunately, we have little information to draw on with longleaf, particularly planted longleaf, and even less information on longleaf planted in old agricultural fields, as is taking place all over the South with the CRP program. The limited data we do have, however, indicates very good growth can be expected if management is done properly and that product yields are very favorable, with a high proportion of poles and quality sawtimber produced.

Dr. David Moorehead, Extension Forester at The University of Georgia s Coastal Plain Experiment Station near Tifton, Georgia has found that on good peanut land in south Georgia and given good survival and quick emergence from the grass stage, longleaf pine will approximate the growth and yield of slash pine, given comparable stocking rates.

We also know that the wood from longleaf is heavier than that of other southern pines. That means that when wood is bought on a weight basis, and it almost always is, more money is paid for longleaf than for the same volume of other pines. One 20 year data set, collected in Mississippi by the consulting firm John Guthrie and Sons, indicates a premium of 10 to 20% paid for sales containing mostly longleaf in every year, in good markets and bad.

In addition, longleaf pine straw has become very valuable in the landscaping business. Returns of $100 to $500 per acre per year have been reported and management techniques for straw production is the subject of much study. One analysis, done by Rick Hamilton of North Carolina State University, predicted a very reasonable internal rate of return for planted longleaf of 7.9% on a site with a site quality index of 45 and an internal rate of return of 9.4% for a site quality of 55. These rates were calculated for revenues earned by sale of wood only. Both are comparable with rates earned by most investments, even the stock market over the long term. When the sale of pine straw is added to the mix, the return rate of the investment increases to

23 9.35% and 10.1% respectively. A general truth in financial analysis is that the earlier in the investment revenues are earned and the later in the investment costs are incurred, the better the investment. This is due to the power of compounding interest and the importance of time when discounting incomes and costs back to the year of investment to make comparisons between investment opportunities possible. Unfortunately, in forestry investments, the opposite is generally the case. Costs are incurred early in the investment and profits are earned later or even at the end of the investment. Early returns from the sale of straw before commercial wood products are produced help longleaf produce income at about the same age as faster starting loblolly or slash. Conservation Reserve Program (CRP) payments offset the early costs of planting very quickly and make forestry and particularly longleaf a very lucrative investment indeed. Since longleaf plantings are currently eligible for CRP contracts of 15 years rather than 10 like other pines, they are particularly attractive. An analysis of longleaf developed by Moorhead and Coleman Dangerfield, a Forest Economist at the University of Georgia, using a planting cost, after cost share, of $97/A, a one-time first year herbicide application cost of $45/A, $10/A per prescribed burn costs at ages 8, 11, and 14, and an annual CRP payment of $40/A yields a very attractive Internal Rate of Return of nearly 29%! Remember, this is after most of the significant costs have been incurred and before the first stick of wood or bale of straw is sold from the land. This return is the result of essentially frontloading the investment with early returns, the CRP payments. This is also an almost entirely risk free investment scenario. The CRP payments are guaranteed by the government if the landowner can keep as few as 200 trees per acre alive during the life of the investment. Longleaf is resistant to diseases and attacks, and notably tolerant of fire, reducing risk of loss to these factors significantly. It is difficult to calculate the value of this risk reduction, but this natural insurance policy against loss does indeed have value. The long term value of this investment is maximized if the trees are allowed to grow into poles, often thought to be optimal in rotations of 55 years or so on most sites, but the CRP payments make it a very profitable investment over the short term as well. Most of us can appreciate the long term value of an investment, say in 50 years, but have a much greater interest in return in terms of our own lifetimes.

Moorhead and Dangerfield estimated Internal Rates of Return of more than 80% using reasonable growth and yield figures, costs, and prices and factoring in aggressive management, wood sales, pine straw, and CRP payments. We cannot predict growth and yield with great confidence at this time, and have even less assurance about things like markets and prices, but it is pretty obvious that longleaf is a good investment.

The Longleaf Alliance is in the planning stages of a regional growth and yield study that should refine our ability to make those projections and make investment analysis more dependable. Other areas of interest include the financial attractiveness of natural regeneration and uneven aged management. The potential for increased wildlife lease value for longleaf plantings is also a subject of much interest. The Longleaf Alliance has adopted the philosophy that the way to save something is to give it value, and one sure way to give something value is to use it. Museums lock up and protect things that once were and wont be again. We don=t want to relegate longleaf to the museum, but make it a contributing component in the southern landscape. In order to accomplish that, it is necessary that private landowners play a significant role. Private landowners have a right to expect a positive return from their lands and on their investments. Longleaf can provide that profit and a host of other benefits as well. All things considered, take a closer look at the economics of planting longleaf pine and PLANT LONGLEAF!

24 MANAGING LONGLEAF FORESTS FOR PROFIT

Barrett B. McCall (RF/ACF President Larson & McGowin, Inc. Forest Managers and Consultants)

INTRODUCTION What constitutes economic success? Should a landowner look to maximize his return on the total capital invested in land and timber (asset value) or look to generate sustained cash flow? Perhaps success simply means having income greater than expense. How do goals such as portfolio diversification come into play?

The typical economic decision criteria employed by most foresters as they look to the management of timberland and the selection of species looks to either growing more wood faster or maximizing return on asset value. Preservation of capital, low risk and inflation hedges are other criteria that are less frequently considered by the analyst but are often extremely important to landowners. Profitable management plans are often over looked or confused when focusing on the maximization of return on asset and/or maximized net present value as the economic criteria. Within industry circles we often hear arguments for natural versus plantation management or loblolly verses all other southern pine species. A landowner can be a profit motivated individual and manage non-loblolly species naturally.

The question then, can you manage a longleaf forest profitably, depends on what constitutes economic success. If maximizing wood fiber production is the goal as it historically has been for many pulp and paper companies, then longleaf in many cases will not be the species of choice. Considering many of the other economic criteria longleaf can provide very good returns and be considered profitable. The following case study along with a few observations will hopefully provide the evidence.

CASE STUDY Inherited in the 1950’s by an attorney from Rhode Island this property of approximately 2,057 acres in South Alabama was managed to generate income and growth and eventually be passed to his five children and eight grandchildren. The several tracts are predominately longleaf pine interspersed with pine hardwood creek branches. Forest practices used on the property have included fire protection, planting of open areas, girdling of cull hardwood and prescribed burning. Income was produced from periodic selective cutting of pulpwood, poles and sawtimber and from hunting lease agreements. As part of a comprehensive estate plan a corporation was formed to hold this and other properties. The last stockholders seldom visited the property, lived in scattered location from Colorado to Maine and preferred managing natural stands to clearcutting and planting. They viewed this holding as an investment diversification, which provided inflation protection, appreciation and some income. The property was divided and sold in September 2000 to several buyers.

The management was volume regulated based on stand conditions, tree size and quality instead of age. Trees typically were 40 to 60 years old at harvest and selectively cut to improve stand conditions including quality and growth. This was accomplished by removing slower growing and poorly formed trees. Care was given to keep tree size and volume stocking balanced in order to prevent high volumes per acre and low stems per acre. The economic results are shown in tables 1 and 2.

Table 1: Average Per Acre Land & Timber Values* Year Land Timber Combined 1943 $3.00 $10.50 $13.50 1958 $27.50 $82.50 $110.00 1976 $300.00 $320.00 $620.00 1992 $436.00 $592.00 $1,028.00 2000 $1,100.00 $1,924.00 $3,024.00

*Does Not Include Income

25 Table 2: Volume and Value Summary* Year Volume Value Starting 1958 2,300 MBF $190,000 Income & Removals 1958-2000 3,600 MBF $1,125,000 Expenses 1958-2000 $265,000 At Final Sale 2000 6,500 MBF $5,970,000 Internal Rate of 1958-2000 10.7% Return

*Volume includes pine sawtimber and poles. Income includes trade of 292 acres.

The ownership certainly for many would be considered a profitable endeavor as demonstrated by the 10.7% return. Income exceeded expense by a sizable margin and the property appreciated in value both in terms of total volume and total value. There were several primary factors that influenced the profits of this longleaf forest including: quality products, low risk (both natural and capital) and stumpage and land price appreciation above inflation as can be seen in Chart 2.

Longleaf naturally produces higher quality sawtimber trees than other southern pines. The merchantable sawtimber volume of a mature longleaf forest is typically 40% to 55% pole quality trees. Loblolly stands by comparison may only contain 5% to 15% pole quality trees. In addition quality sawtimber will typically be worth between 10% to 25% more than average sawtimber. Chart 1 illustrates the relationship between small sawtimber (chipnsaw), sawtimber and poles.

The management of this forest took advantage of the low risk characteristics of longleaf forests. As many recognize the longleaf pine is naturally fire, drought, insect and disease resistant. In addition longleaf lends itself to natural management systems that take advantage of low capital risk. By managing this forest with selective harvest keeping the inventory weighted towards older larger sawtimber sized trees increased liquidity of the standing inventory. This increased liquidity provided market flexibility. If markets were down the landowner didn’t sell. If markets were up the landowner took advantage of higher prices and harvested timber. The natural characteristics of longleaf allow the landowner to be patient about the timing of harvest without too much concern of natural loss to insect and disease. This is true for other southern pines but more so with longleaf.

This forest was managed at low cost thus not requiring extra capital investment from the landowner. Managing a forest “extensively” can be very profitable but requires expertise. Often the decision of the person with the paint gun selecting the trees to harvest can make or break the success of low cost management. Constant consideration should be given promoting regeneration without promoting its competition. Keeping watch over seed crop production and carefully scheduling burning can also effect the success of natural regeneration.

26

Chart 1: Southwest Louisiana Pine Stumpage Prices $$/Ton

$80.00

$70.00

$60.00

$50.00

$40.00

$30.00

$20.00

$10.00

$- 1997 1998 1999 2000

Pine Poles Pine Sawtimber Pine Chipnsaw Pine Pulpwood

Source: 70 Larson & McGowin, Inc. Manages Sales

27 Chart 2: Average Land & Timber Prices South Alabama

$900

$800

$700

$600

$500

$400

$300

$200

$100

$0 1926 1936 1946 1956 1966 1976 1986 1996

Land/Acre Pine Sawtimber/MBF Doyle Pine Pulpwood/Cord

Source: TimberMart South and Larson & McGowin, Inc. Land & Timber Sale Database

CONCLUSIONS Longleaf can be managed profitably. Its growth and economic return characteristics are different from other southern pine species but often match the economic goals of the landowner. With the decline in the pulp and paper markets in the south over the last few years the case for longleaf gets stronger. Landowners will likely become less reliant on chipnsaw and pulpwood as their end products and emphasis will be placed back on growing higher valued products like sawtimber and poles. In this situation, when one evaluates growing value verses volume, longleaf is very attractive. To be successful the landowner should utilize all of the silvicultural tools available and not fall victim to the single species single treatment methods. Consider the balance between expense of money and the expense of time. When objectives call for quick results or stand conditions are unfavorable usually capital expenditure is necessary. If time is available and activities are well planned then low cost management will succeed.

28 NATURAL STAND DYNAMICS IN LONGLEAF PINE: HOW CLIMATIC DISTURBANCES SHAPE THE COMMUNITY

Kenneth W. Outcalt (USDA Forest Service, Athens, GA 30602)

INTRODUCTION Longleaf pine (Pinus palustris) once dominated the overstory of a wide range of southern plant communities including wet flatwoods and savannas of the Atlantic and Gulf coastal plains to higher, droughty sand deposits such as the fall line sandhills and the central ridge of Florida. It was also the dominant overstory tree on the lower rolling sandhills of lower Mississippi, central Louisiana, and east Texas. Longleaf pine even extended onto the mountain slopes and ridges of Alabama and northwest Georgia, where it was found growing at elevations up to 600 m. Frequent low intensity fires favored longleaf over other pines, because it has seedlings better adapted to surviving such fires. These periodic fires, every 2 to 8 years, also shaped the rest of the community, keeping woody shrubs in check while promoting the growth of grasses and forbs.

STAND LEVEL DISTURBANCE Although fire is a primary driving force in longleaf pine ecosystems, it is not the only disturbance that impacts their structure and composition. A good portion of the longleaf range lies within 100 miles of the coast. Thus, stands have a rather high probability of being impacted by a significant hurricane at least once every 100 years (Hooper and McAdie 1996). During strong hurricanes like Hugo, which hit South Carolina in 1989, the tallest and largest trees receive the most damage and are usually uprooted or broken off (Hook and others 1996). This leaves behind the less severely damaged midstory trees. This scenario creates a sort of natural clear-cut of the overstory with a relatively even-sized sapling and seedling stand remaining. Prior to settlement, the tremendous fuel loadings would have greatly increased the probability of severe wildfires following these hurricanes. With or without such fires, the net result was an open area that could be occupied with longleaf pine regeneration either from existing individuals or by establishment from seeds shed by trees which escaped the fires. The even-aged management system, long used for longleaf pine, mimics these natural disturbance and regeneration patterns.

Because of the numerous thunderstorms that occur in the South, tornadoes are also prevalent within the longleaf range. In South Carolina for example, there are about 10 tornados per year (Purvis 1990). The most destructive are those, which remain on the ground for many miles. Tornadoes leave a path of broken and twisted dead overstory trees. Like severe hurricanes, this creates an open area that is captured by pre-existing longleaf seedlings or colonized by new seedlings from seed shed by adjoining trees. This results in a relatively even sized, although not entirely even-aged, stand of longleaf pine.

SMALL SCALE DISTURBANCE The importance of lightning as an ignition source for the fire driven longleaf pine ecosystem is widely recognized. Lightning also impacts this system on a smaller scale by causing individual tree mortality. Often small groups of 2 to 4 trees are killed, creating small gaps in the longleaf canopy. Gaps are also created through the cumulative effect of individual tree mortality over multiple storms and years. Thunderstorm activity, and therefore lightning strikes, varies with location across the south. I followed lightning activity for 4 years in longleaf stands on the Department of Energy’s Savannah River Site located in west central South Carolina. A total of eight stands at three locations containing 255 ha were monitored for lightning mortality. Lightning directly or indirectly killed 1 tree/6ha/year. A similar study on the Ocala National Forest in central Florida, the lightning capital of the world, had double the rate at 1 tree/3ha/year. Over a 100-year period, lightning mortality would remove 5% of the overstory in the South Carolina stands and 12.5% of the overstory in the Florida stands. This research also showed that it selectively removes the largest trees in the stand.

On sandhills sites, like the Ocala National Forest, longleaf pine regeneration is concentrated in these lightning created gaps. Regeneration, however, is not uniform across gaps. Brockway and Outcalt (1998) showed seedling density was much higher in the center of the gaps with a significant increase beyond 12m from the gap edge, where the gap edge was defined as the bole of trees surrounding the gap rather than the crown drip line. This distribution of seedlings was negatively correlated with fine root and forest floor

29 biomass. In subsequent research, on the Ocala National Forest, I looked at natural seedling establishment and survival across gaps with diameters of 30 to 40m. Seedling establishment was uniform from the edge to the center of gaps. Survival, however, was significantly lower for seedlings less than 4m from the edge, i.e. under the crowns of edge trees. This was correlated with rainfall distribution within the gaps. The first prescribed burn at the end of the second growing season reduced survival at all locations to less than 5%. A second fire, 3 years later, further reduced seedling survival and eliminated differences between locations.

In another study in the same longleaf gaps, I established containerized seedlings at 1, 4, 8, 12, and 16 meters from gap edges. At each location, seedlings were planted in a control patch and inside root control rings 0.5m in diameter and 20cm wide installed so the top was at the soil surface. Mean survival was higher for seedlings planted in rings that reduced root competition. The effect disappeared beyond 12m and from there to the center of the gap, reducing root competition had no effect on survival. This corresponded to the distance where fine root biomass declined in the study by Brockway and Outcalt (1998). A prescribed fire 2 years after planting reduced seedling survival at all locations, but the reduction was greatest under the crowns of edge trees. With reduced root competition, survival of longleaf seedlings after 5 years, with a prescribed fire at age 2, was just as good at 4m as in the center of gaps.

Thus, a number of interacting factors cause the distinctive distribution of longleaf seedlings and saplings in gaps of droughty sandhills sites. There is a zone from 0 to 4 or 5m on the gap edge where seedling establishment is equal to that found throughout the gap, but first year survival is lower. This is caused by interception of rainfall by tree crowns and likely poorer moisture holding capacity of the thicker forest floor. There exists a second zone that spans the area from 0 to 12m where fine root competition, likely for moisture, reduces survival of older seedlings. The frequent fire that occurs in longleaf stands also interacts with differences across the gaps. Seedlings in the 0 to 4m zone are more susceptible to fire mortality because the fuel loads are greatest under the crowns of the edge trees. They are also more susceptible, as are the seedlings in the 4 to 12m zone, because root competition reduces growth rates. This means it takes longer for the seedlings to reach a root collar diameter where they are likely to survive fire. Thus, even if a seedling in the 0 to 12m zone does survive the droughty conditions, it will likely be killed by fire because of its slow rate of growth. Beyond this 12m exclusionary zone is the central area of the gap where seedlings have a higher probability of survival.

CONCLUSIONS Climatic caused disturbances significantly impact longleaf pine communities changing stand structure and providing open sites for regeneration. Severe hurricanes operate at the landscape scale creating a type of natural clear-cut by removal of most of the large overstory trees. Tornadoes, which usually operate at the partial stand scale, also create open conditions where regeneration can occur. Even-aged management of longleaf pine mimics these natural disturbance and regeneration patterns. Seed-tree and shelterwood systems create conditions similar to less severe hurricanes that remove only some of the overstory. Lightning strikes, although they affect less area in a stand, are continuously impacting longleaf stands creating small-scale gaps of 2 to 4 trees. Regeneration in the small gaps is not uniform because of variation in precipitation, litterfall, and root distribution and their interaction with frequent fire. Managers using the selection system should be aware of this, and create gaps in dry sandhills sites accordingly. The minimum size is about 0.08 hectare or a radius of 16m. The ideal size, however, is about 0.2 hectare or a gap with a radius of 25m, because one gap that size has 27% of its area in the zone where seedlings are likely to survive versus just 12% for two of the small gaps. If gaps larger than 0.2 hectares are desired, they should be oblong in shape with a maximum width of 50m so seeds do not have to travel more than twice the height of seed trees.

LITERATURE CITED Brockway, D.G., and K.W. Outcalt. 1998. Gap-phase regeneration in longleaf pine wiregrass ecosystems. Forest Ecology and Management 106:125-139. Hook, D.D., M.A. Buford, and T.M. Williams. 1996. Impact of hurrican Hugo on the South Carolina Coastal Plain forest. In J.L. Hammond, D.D. Hook, and W.R. Harms (eds), Hurricane Hugo: South Carolina Forest Land Research and Management Related to the Storm. U.S. Department of Agriculture, General Technical Report SRS-5, Southern Research Station, Asheville, NC. Pp.34-43.

30 Hooper, R.G., and C.J. McAdie. 1996. Hurricanes and the long-term management of the red-cockaded woodpecker. . In J.L. Hammond, D.D. Hook, and W.R. Harms (eds), Hurricane Hugo: South Carolina Forest Land Research and Management Related to the Storm. U.S. Department of Agriculture, General Technical Report SRS-5, Southern Research Station, Asheville, NC. Pp.417-436. Purvis, J.C. 1990. South Carolina tornado statistics. Southeast Regional Climate Center, 59p.

31 LONGLEAF AND THE PRIVATE, NON-INDUSTRIAL LANDOWNER OUTSTANDING OPPORTUNITIES

Michael A. Webb, A.C.F. (Webb Forestry Consultants)

Small private non-industrial landowners (P.N.I.L.*), represent one of the greatest opportunities for restoration of Longleaf. Not only do they own around 65 % of the original Longleaf lands, but collectively have hundreds of millions of dollars to invest in Longleaf restoration and management. Our responsibility is to show them how Longleaf Management is best suited to meet their needs.

In order to do so, our first step then is to understand these landowners, their situation, needs and goals. Obviously while perspectives and goals will vary significantly among landowners, there are certain generalizations that can be made. Table 1 summarizes some generalizations about the perspective of the P.N.I.L. While some are fairly clear, some explanation is important.

- Shorter Outlook - Whereas Longleaf can live for 300 years, most Landowners are looking for a return within their lifetime. For many people their land is their most valuable asset. It is their retirement income, their children's college fund, and/or a source of additional income if not their outright livelihood. It is their estate to pass on to their children.

Therefore, a true sensitivity has to be kept in mind of the landowner's point-of-view in helping him or her see the opportunities for Longleaf to fulfill those financial needs, while at the same time allowing them to be a steward of the land but not feel threatened for caring for the longleaf and the attendant ecology that goes with it.

- Educational Needs - Most private landowners do not understand forestry in general, and Longleaf in particular, to know what is best to meet their own objectives. It is estimated that less than 30% even have a written plan for their forest lands.

*I am specific to say "Landowner" as opposed to forest landowner, as many acres of the original Longleaf lands are now in pasture or row crop that has excellent potential for Longleaf restoration but are owned by landowners whom do not see themselves as forest landowners.

- Smaller Acreage - Whereas the original Longleaf forest is pictured as once having been a vast unbroken expanse, the reality today of severely broken up ownership. Even leaving out the 5 and 10- acre ranchettes, the average ownership is still under 150 acres.

- Not Longleaf - Of the historically "Longleaf" lands, only a small fraction is in Longleaf today and this is especially true of the private ownership, with most in pasture, loblolly or slash plantations, or mixed hardwood-pine stands.

- Anxiety About Operations - Obviously, as the land means so much to private landowners, and they will be investing in management they are very concerned about how successful management is going to be. The enormous amount of misinformation about Longleaf and the necessity for special tools such as prescribed burning require that we continuously work with P.N.I.L. to help them understand Longleaf Management.

- Diversity of Goals - Particularly unique to private landowners is the diversity of goals for ownership not only between landowners, but also for the same landowner. Table #II lists the goals that landowners often give for ownership; and generally most landowners will see two or three, if not all of these of some importance.

Further, many studies have surprisingly found income and return on investment often listed lower on landowner's hierarchy of goals, with aesthetics, recreation, and wildlife being higher. Even so, we do have to

32 keep in mind, when landowners have a need for money or an opportunity to receive $150,000.00 - $200,000.00 is suddenly put forward, these priorities sometimes take an unexpected reversal. While these goals are somewhat self-explanatory, a couple does need further definition.

-Heritage - so often reflected in land inherited and a desire to pass it on to future generations.

- Investment - While some buy land purely as an investment and keep it only so long as it can exceed the potential return of other investments; most landowners do not see a maximum return as so critical and are willing to give up some return for other values. At the same time, being able to show a good investment value is an extremely valuable tool for encouraging landowners to invest in Longleaf restoration and management.

Having now briefly looked at some of the major factors influencing private landowners, our next step is to then consider the values of Longleaf and Longleaf Management to meet these goals for the private landowners.

In looking over the list of landowner goals, the fact is Longleaf should be an easy sale. The advantages of Longleaf for aesthetics, recreation, and wildlife are obvious. Yet, perhaps one of our best selling advantages is for protection since Longleaf is, by far, the most resistant of all the Southern pines to every major danger of the Southern Pine Forest, be it wildfire, pine beetle, disease or storms.

Heritage as a goal, also is a strong value for Longleaf, and helps us overcome the "shorter outlook" as it involves generations, which can allow for longer rotations. Further, these landowners tend to put more significance on other values besides economics.

In like manner, while smaller acreage is in some ways a disadvantage; it can also allow an affordable pursuit for someone to rebuild a natural heritage as Author, Don Schueler, has done and documented in his book "A Handmade Wilderness".

Given that the major tools of Longleaf Management of prescribed burns, selective harvesting, and natural regeneration, as well as the species itself, being more superior for achieving the aesthetics, recreation, wildlife and protection that landowners see as major goals of ownership, why hasn't Longleaf been more successful in recent years?

Historically, there are a number of reasons for the failure of our Longleaf forests to come back after the wholesale cuts of the early 1900's, including the loss of natural fire to control brush and grass competition to young Longleaf seedlings, and the poor survival and growth of early Longleaf plantings being most prominent. However, the main reason, over the last twenty years has simply been a lack of education of foresters, county agents and landowners that new techniques have resolved these problems.

The word has not gotten out that Longleaf can be reforested successfully, that height growth can be initiated quickly and economic returns are very good if not perfectly comparable to Loblolly plantings. Which unfortunately, is still almost automatically the species of choice for reforestation among private landowners and foresters.

Finally, with only two to five percent of the original Longleaf range still supporting adequately stocked stands of Longleaf, while management of this residual standing Longleaf is very important, critical efforts need to focus on reforestation back to Longleaf of both old fields as well as cutover sites.

CONCLUSIONS

1.) Small Non-Industrial Private Landowners own the majority of the original Longleaf lands and will be the key to significant restoration of Longleaf. 2.) Based on an understanding of the private landowners, their needs and goals, an overall plan needs to be developed to reach these private landowners and help them understand the significance and value of restoring and managing Longleaf Pine.

33 3.) On most sites, Longleaf Pine is optimally suited to meeting the goals of private Non-Industrial Landowners. 4.) Because of the management focus of prescribed burning, selective harvests and natural regeneration, it is particularly valuable for achieving major P.N.I.L. goals including aesthetics, recreation, wildlife and safety. 5.) Due to technological advances, Longleaf is now well suited for reforestation and capable of producing excellent investment returns. 6.) Establishing educational programs about the opportunities of Longleaf, first for County Foresters and County Agents, then foresters in general and finally landowners in general is the critical step in beginning the restoration of the Majestic Southern Longleaf Forests.

TABLE #I

Generalizations about the Private Non-Industrial Landowners (P.N.I.L.)

-Shorter Outlook -Educational Needs -Smaller Acreage -Income Needs -Major Not Presently in Longleaf -Anxiety about Operations -Diversity of Goals

TABLE #II

Landowner Goals

-Recreation -Aesthetics -Heritage -Timber Production -Wildlife -Income -Return on Investment -Protection

34 ECOLOGY AND MANAGEMENT OF GROUND COVER IN LONGLEAF PINE FORESTS OF THE WEST GULF COASTAL PLAIN

Rick L. Turner (The Nature Conservancy, 506 Hayter St., Nacogdoches, TX 75961)

INTRODUCTION Longleaf pine forests are widely recognized as one of the most diverse terrestrial plant communities in North America. However, the overstory of these forests is usually dominated by one species, longleaf pine, and perhaps a few other associated tree species. Virtually all plant diversity in longleaf pine ecosystems is present in the ground cover layer. Ground cover can be defined as annual and perennial herbaceous plants and low- growing woody shrubs that are generally less than 1 meter in height at maturity. Ground cover not only contains the vast majority of the biological diversity in longleaf pine forests, but if managed properly it can also contribute economic benefits to landowners and managers. The following discussion will address the ecological and economic values of ground cover in longleaf forests of the West Gulf Coastal Plain, as well as management issues related to maintaining a diverse and abundant groundcover.

CLASSIFICATION AND ECOLOGY Several classifications have described longleaf pine communities of the West Gulf Coastal Plain at various levels of scale, using both quantitative and qualitative methods (Watson 1975, Bridges and Orzell 1989, Martin and Smith 1991, Harcombe et al. 1993, Weakley et al. 1999, Turner 1999, Turner et al. 1999). For the purpose of brevity, I will describe three general types of longleaf pine ecosystems and associated ground cover that occur in the West Gulf Coastal Plain ecoregion: dry uplands, mesic uplands, and wet flatwoods.

Dry Uplands occur mainly on sandy, well-drained soils on ridgetops, upper slopes, and some sandy stream terraces. This community type tends to be more common in areas of the region with greater topographic relief. Longleaf pine may dominate natural stands, but it often occurs in mixed stands with shortleaf pine (Pinus echinata), sand post oak (Quercus margarettiae), bluejack oak (Quercus incana), and black hickory (Carya texana). Some steep, isolated ridgetops may be dominated by oaks and hickories. Ground cover plant species are tolerant of periodic drought during the growing season, and they also exhibit adaptations to a frequent fire regime. Characteristic species include bullnettle (Cnidoscolus texana), noseburn (Tragia urens), longleaf buckwheat (Eriogonum longifolium), hairy bush clover (Lespedeza hirta), soft greeneyes (Berlandiera pumila), spiderwort (Tradescantia reverchonii), queensdelight (Stillingia sylvatica), Louisiana yucca (Yucca louisianensis), and pricklypear cactus (Opuntia spp.).

Mesic Uplands occur on finer textured, moderately well-drained, loamy soils. On these sites, soil moisture is more abundant than on the sandy dry uplands. Longleaf pine is almost always dominant in the overstory and it often forms pure, dense stands. Hardwoods that may be present include post oak (Quercus stellata), southern red oak (Quercus falcata), black gum (Nyssa sylvatica), and flowering dogwood (Cornus florida). Due to favorable growing conditions, ground cover growth is often luxuriant, supporting a highly diverse assortment of species. Little bluestem (Schizachyrium scoparium), the typical grass species of most longleaf pine communities of the West Gulf Coastal Plain, is often found in greatest abundance on mesic upland sites. Other characteristic species of mesic uplands include pencil flower (Stylosanthes biflora), silkgrass (Pityopsis graminifolia), Virginia hoarypea (Tephrosia virginiana), New Jersey tea (Ceanothus americanus), and purple coneflower (Echinacea pallida).

Wet Flatwoods are found on nearly level plains with poorly drained, nutrient-poor soils and a seasonally high water table. The overstory is often sparse and is usually dominated by longleaf pine, sometimes with a sparse midstory of swamp blackgum (Nyssa biflora), sweetbay magnolia (Magnolia virginiana), and other woody species tolerant of saturated soils. The species-rich ground cover is usually dominated by native grasses and sedges, especially beakrushes (Rhynchospora spp.). Other indicator species include tenangle pipewort (Eriocaulon decangulare), blueflower eryngo (Eryngium integrifolium), swamp sunflower (Helianthus angustifolius), and pinewoods rosegentian (Sabatia gentianoides).

Within the three broad longleaf pine community types are imbedded distinctive plant communities that are associated with particular environmental conditions. Hillside Seepage Bogs occur on upland slopes where groundwater seeps to the surface, creating locally wet, acidic, nutrient poor soils. Species composition is

35 similar to that of the wet flatwoods, except that the ground cover may be dominated by insectivorous plant species such as yellow pitcher plants (Sarracenia alata). Baygalls and Bayheads occur in shallow, constantly saturated drainages associated with small streams. They support an abundant growth of woody vegetation, including sweetbay magnolia, possumhaw (Viburnum nudum), redbay (Persea borbonia), and gallberry holly (Ilex coriacea). The ground cover is often dominated by various fern species. Glades and Barrens are prairie- like openings in the forest that occur on rocky outcrops, high shrink-swell clays, and calcareous or saline soils. Species composition can vary by site, and several rare or disjunct herbaceous species occur only in these communities. Stream Bottoms and Terraces are hardwood-dominated, closed-canopied forests with sparse groundcover, located along the lower slopes or floodplains of small streams. Depression Marshes are low flooded areas within flatwoods and are typically dominated by floating and rooted herbaceous aquatic vegetation.

Soil texture, moisture content, and nutrient availability can greatly influence the species composition of ground cover. Soil characteristics largely depend on the composition of underlying bedrock from which it developed. In the West Gulf Coastal Plain, geologic formations generally occur in bands that roughly parallel the present shoreline of the Gulf of Mexico. The age and topographic relief of theses formations generally increases from south to north in the region.Almost all of the formations are sedimentary and were formed from deposits of either continental or marine origin. The chemical and physical characteristics of these deposits will in turn affect the types of soil that develop.

On most upland landscapes in the West Gulf Coastal Plain, frequent, low-intensity fire is the most important ecological process that maintains herbaceous ground cover in longleaf pine forests. Naturally ignited fires normally occur during summer, when thunderstorm and lightning activity is highest and seasonal droughts are common (Komarek 1968). Virtually all native ground cover species (as well as longleaf pine seedlings) have adaptations or life histories that favor them in a fire-dominated environment. Most species are perennials which are readily able to regenerate from rootstock after their above-ground portion has been consumed by fire.

MANAGEMENT A diverse, intact ground cover has potential to provide several economic and management benefits. Ground cover provides fine fuels for frequent low-intensity fires necessary to maintain and naturally regenerate longleaf pine forests. Frequent, low-intensity fires also greatly reduces the risk of tree loss from wildfires. An open understory improves the efficiency of cruising, timber marking, and tree harvesting activities. A diverse ground cover provides excellent habitat for game species such as bobwhite quail, wild turkey, and white- tailed deer, allowing opportunities for improving hunting leases. It also enhances the scenic value of longleaf pine forests by supplying an interesting diversity of plant species and non-game wildlife, such as songbirds and butterflies. Recreational users including hikers, campers, and birdwatchers benefit from easier travel and increased visibility.

Management for groundcover is an important part of restoring and managing longleaf pine forests in an ecologically sustainable way. Since fire is such an important process in longleaf pine ecosystems, it should be part of any management program. Natural fire frequency varies by climate and local site conditions (generally between 1 and 10 years). Due to fewer natural firebreaks, wet flatwoods burn more frequently than topographically isolated upland ridges. On most landscapes where a natural fire is unlikely to occur or is undesirable, prescribed fire must be used. Growing season burns are best for controlling understory shrubs and trees, since much of the carbohydrates stored in roots have been used to produce new growth and have not had a chance to be replenished (Boyer 1990, Rebertus et al. 1993). While season of burn has not been conclusively shown to significantly affect ground cover composition, growing season burns approximate the natural fire regime to which most native plant species are adapted (Streng et al. 1993).

Herbicides have sometimes been used instead of fire to control competing vegetation in pine plantations. Not much is known about the effects of herbicides on many ground cover plants, especially rare species. Many species require fire at a certain time in the growing season to stimulate flowering and seed production, so the lack of periodic fire could reduce reproduction even if the plants are not directly harmed by herbicides. On some restoration sites that have been subjected to long-term fire suppression, herbicides have been used in spot applications on hardwood growth that is too large to be killed by frequent fires alone.

36 Little information exists on the short-term effects of pine straw raking on ground cover diversity, and long- term effects are essentially unknown. One study found that raking did not increase non-native plants and generally had little effect on species composition (Kelly et. al 2000). Most ground cover species in longleaf pine forests are perennials and are adapted to periodic above-ground destruction by fire. However, frequent rakings could contribute to long-term shifts in composition by reduction of fuels for frequent fires, loss of fertility, and damage to longleaf pine reproduction.

A thorough discussion of various silvicultural techniques for managing timber in longleaf pine forests and their effects on ground cover diversity is beyond the scope of this paper. Farrar (1996) provides a practical discussion of the advantages and disadvantages of uneven-aged management of southern pine forests. One potential problem with even-aged management is that prescriptions often call for intense site preparation before regenerating the stand. Some ground cover species are slow to re-colonize after soils are severely disturbed, so any kind of site preparation other than fire may eventually lead to extirpation of rare or shallow- rooted species (Mejeur et al. 2000). Whether even-aged or uneven-aged management is selected, the silvicultural options that will be most successful in maintaining ground cover diversity are those that include three essential elements: minimal soil disturbance, a relatively open tree canopy, and frequent prescribed fires.

CONCLUSIONS Ground cover is an essential part of the species diversity of longleaf pine ecosystems. Not only does ground cover support important ecological processes such as fire, it also can provide added economic value. Each of the three general types of longleaf pine communities occur on the West Gulf Coastal Plain--dry uplands, mesic uplands, and wet flatwoods--has its own characteristic set of ground cover species. Within these general types are found many specialized inclusional communities, including hillside seepage bogs, baygalls and bayheads, glades and barrens, and depression marshes. Species composition in longleaf pine communities is influenced by multiple environmental factors, especially fire frequency and soil/topographic conditions. Maintaining ground cover diversity in managed longleaf pine stands is possible if conditions favorable to ground cover, namely an open overstory canopy, minimal soil disturbance, and frequent fire, are provided.

LITERATURE CITED Boyer, W.D. 1990. Growing season burns for control of hardwoods in longleaf pine stands. USDA Forest Service, Southern Forest Experiment Station Research Paper SO-256. Bridges, E. L., and S. L. Orzell. 1989. Longleaf pine communities of the West Gulf Coastal Plain. Natural Areas Journal 9:246-263. Farrar, R.M. 1996. Fundamentals of uneven-aged management in southern pine. Tall Timbers Research Station Miscellaneous Publication No. 9, Tallahassee, FL. Harcombe, P. A., J. S. Glitzenstein, R. G. Knox, S. L. Orzell, and E. L. Bridges. 1993. Vegetation of the longleaf pine region of the West Gulf Coastal Plain. Proceedings of the Tall Timbers Fire Ecology Conference 18:83-104. Kelly, L.A., T.R. Wentworth, and C. Brownie. 2000. Short-term effects of pine straw raking on plant species richness and composition of longleaf pine communities. Forest Ecology and Management 127:233-247. Komarek, E.V. Lightning and lightning fires as ecological forces. Proceedings of the Tall Timbers Fire Ecology Conference 8:169-197. Marks, P. L., and P. A. Harcombe. 1975. Community diversity of coastal plain forests in southern East Texas. Ecology 56:1004-1008. Mejeur, R.S., J.L. Walker, and B.P. Van Eerden. 2000. Herbaceous diversity and spatial pattern in longleaf pine communities: a comparison of natural stands and plantations (abstract). Tall Timbers Fire Ecology Proceedings 21:40. Rebertus A.J., G.B. Williamson, and W.J. Platt. 1993. Impact of temporal variation in fire regime on savanna oaks and pines. Proceedings of the Tall Timbers Fire Ecology Conference 18:215-225. Streng, D.R., J.S. Glitzenstein, and W.J. Platt. 1993. Evaluating effects of season of burn in longleaf pine forests: a critical literature review and some results from an ongoing long-term study. Proceedings of the Tall Timbers Fire Ecology Conference 18:227-263

37 Turner, R.L. 1999. Ecosystem classification of four national forests on the West Gulf Coastal Plain of Texas. MS Thesis, Stephen F. Austin State University, Nacogdoches, TX. Turner, R.L., J.E. Van Kley, R.E. Evans, and L.S. Smith. 1999. Ecological classification system for the national forests and adjacent areas of the West Gulf Coastal Plain. Report to the U.S. Forest Service, The Nature Conservancy, Nacogdoches, TX. Watson, G. 1975. Big Thicket plant ecology: an introduction, 2nd edition. Big Thicket Museum Publication Series No. 5, Saratoga, TX. Weakley, A. S., K. D. Patterson, S. Landaal, M. Pyne, M. J. Russo, and others (compilers). 1999. International classification of ecological communities: Terrestrial vegetation of the southeastern United States. The Nature Conservancy, Chapel Hill, NC.

38 “LESSONS IN LONGLEAF”: OBSERVATIONS FROM THE NORTH CAROLINA SANDHILLS

Terry Sharpe (NC Wildlife Resources Commission)

The bobwhite quail is a widely recognized member of the suite of grassland/early succession birds, which have undergone widespread dramatic population declines since the mid-1960s. Data from hunter check stations and call-count routes indicate that similar bobwhite population declines have occurred on Sandhills Game Land and Ft. Bragg Military Reservation in the North Carolina Sandhills. These large public holdings do not have a history of forest conversion, agriculture, or pesticide use often linked to bobwhite declines in other areas. Dominant upland vegetation types (longleaf, turkey oak, and wiregrass) have not changed.

Recent (1988) and old (Circa. 1950) aerial photography of Sandhills Game Land was compared to photography of impact areas on Ft. Bragg which are subjected to almost annual growing season fires and still support high bobwhite populations. Comparisons reveal that bobwhite population declines on Sandhills Game Land and Ft. Bragg can be linked to increased tree canopy from pines and hardwoods, particularly in wetlands and along fertile upland/wetland transition zones. Prior to the 1970s extensive logging and frequent wildfires which burned across sandhills drains maintained grassland and savanna habitats in a diverse herbaceous community. Due to wildfire suppression, burning uplands with cool winter fires, and more conservative timber harvests the open canopy, rich herbaceous communities which support high bobwhite populations now exist in substantial acreage only on impact areas on Ft. Bragg where they are maintained by frequent fires set by ordinance.

A series of 22 annual growing season burns on one small wetland complex on Sandhills Game Land has created and maintained a diverse herbaceous plant community. This area provides an example of the attributes that an open forest canopy and diverse herbaceous plant community provide. Observations and a series of short-term research projects on the area have documented that the “diversity (species richness) in both vegetation and seed reserves, in all moisture regimes, increased with an increase in the number of years an area had been burned…..” (May 1981). Sweep net samples yielded twice as many on the annual burn as found on adjacent areas winter burned every three years with winter fires. Tall native warm season grasses provide concealment. Patches of grassy and brushy cover remain in wetlands after annual burning due to variations in fuel loads, moisture, and wind conditions. Other sandhills wildlife species benefiting from the open forest stands, cover conditions, and food resources provided by annual burning include: Red Cockaded Woodpeckers, Bachman’s Sparrows, and numerous herps.

Land managers should recognize the importance of diverse groundcover to wildlife and ecosystem functions. Restoring and maintaining the herb layer in fertile transition zones and wetland communities is a challenge to land managers, but is critical to the restoration of bobwhites and other wildlife dependent upon the frequent fire regimes typical of the longleaf ecosystem.

LITERATURE CITED May, M.L. 1981. Vegetation, seed reserve, and environmental relationships in the sandhills. Unpublished MS Thesis, UNC Chapel Hill.

39 USING SAFE HARBOR AGREEMENTS TO BENEFIT LANDOWNERS AND THE LONGLEAF PINE ECOSYSTEM

Lee Andrews (U.S. Fish and Wildlife Service, Southeast Region, 1875 Century Blv., Suite 200, Atlanta, GA 30345)

SAFE HARBOR PROGRAM BACKGROUND In the early 1990's, the North Carolina Sandhills Region was in turmoil over the most basic Endangered Species Act (ESA) issue: How do you balance the needs of private landowners with the needs of federally listed species without causing undue harm to both? The Fish and Wildlife Service (FWS) had been contacted by a landowner who wanted to remove hardwoods from a mature longleaf pine stand, thin the longleaf stand, and initiate a prescribed burning schedule in order to improve bobwhite quail habitat. The trouble was that doing so would almost certainly encourage the use of this longleaf stand by the endangered red-cockaded woodpecker (Picoides borealis) (RCW), which were known to live nearby but not within the longleaf stand in question. The landowner wanted to improve the habitat on his land but did not want to incur any additional ESA responsibilities for doing something that would have a positive impact on RCWs. Couldn’t the FWS do something to help the RCW without putting a regulatory burden on the landowner? Surely, there were other landowners in the same predicament.

Unsurprisingly, there were other landowners in the same predicament, and, once the FWS started looking, similar situations sprang up all over the country. From other North Carolina landowners with RCW issues to Texas ranchers worried about listed songbirds, the response was the same. Landowners were, in many instances, interested in helping with listed species management but only if they would not incur additional ESA regulatory liability, regardless of whether that liability was perceived or real. What the landowners were asking for was land management flexibility in spite of listed species, in a sense, “safe harbor” from future land use restrictions if they provided benefits to listed species. In response to landowner demand, the FWS began working on a Safe Harbor policy and implementation procedure that was finalized in June 1999 (Federal Register 1999a,b).

WHAT DOES SAFE HARBOR DO? Some landowners are hesitant to manage their lands for listed species, to restore habitat for listed species, or to allow increases in the populations of listed species on their lands. This hesitance stems from the fear that the FWS would impose additional restrictions on the use of the land. SHAs address an important need common to private forest landowners who own property that contains (or could contain) listed species or their habitat. That is, SHAs provide landowners regulatory assurances under the ESA that help provide the management flexibility most landowners want. With these regulatory incentives in place, many landowners are willing to voluntarily manage their property for listed species. Generally, these property owners will agree to restore, enhance, or maintain habitat for listed species or reintroduce the species in return for the FWS’s assurance that additional land, water, or other use restrictions will not be imposed as a result of the landowner’s voluntary conservation actions.

In signing a SHA, the FWS agrees to not penalize landowners who willingly provide management for listed species. Using these assurances as the foundation to build trust between the FWS and landowners, the Safe Harbor program encourages the cooperation of private landowners in management programs for listed species. The cooperation of private landowners is imperative, because conservation and recovery of many listed species is primarily a private lands issue. Many listed species occur exclusively, or to a large extent, on non-federal land.

INCENTIVES TO LANDOWNERS The Safe Harbor program can provide three primary incentives to participating landowners. First, it can provide a land management incentive. When landowners are not worrying about the increased or future presence of listed species, they can manage the land with greater flexibility. A SHA can allow this to happen, as long as no existing listed species or occupied habitat are harmed. Second, it can provide an economic incentive, such as encouraging landowners to employ longer timber rotations to diversify their timber revenue sources, or encouraging landowners to improve habitat conditions by thinning or burning in order to increase hunting-related revenues. Third, it can provide a regulatory incentive. Section 9 of the

40 ESA prohibits the “take” of a listed species, and “take” is defined to include such actions as killing, pursuing, harassing, or harming, which includes modification or destruction of necessary listed species habitat. Landowners would be in violation of the ESA if they caused the “take” of a listed species. However, section 10 of the ESA allows the FWS to permit certain types of take as long as that take is incidental to an otherwise lawful activity, such as timber harvesting, ranching, or residential development. For Safe Harbor programs, the FWS issues Enhancement of Survival Permits (ESPs) under section 10(a)(1)(A) of the ESA. ESPs provide landowners the authority to reduce the habitat or populations of a listed species back to the original (i.e., baseline) level. Issuance of an ESP also ensures that additional conservation measures and land use restrictions will not be required of the landowner, even if the species becomes more numerous or its habitat increases. After all, these individuals or their habitats would not have existed without the landowner’s voluntary conservation actions.

HOW SAFE HARBOR PROGRAMS WORK To enroll in the Safe Harbor program, a landowner works with the FWS to develop a Safe Harbor Agreement (SHA) that specifies the conservation action(s) that the landowner intends to implement, the schedule that the conservation actions will follow, and the baseline conditions present on the property. Identification of accurate baseline conditions is crucial to providing regulatory assurances to the landowner. Establishing baseline conditions usually involves a survey of the property and mapping of available or potential listed species habitat, and, once this is done, the baseline is usually expressed in terms of the number of individuals or amount of habitat present at the time the SHA is signed.

In order to finalize a SHA, the FWS must make a determination that the proposed SHA would provide a “net conservation benefit” to the listed species. The Safe Harbor policy is designed to provide landowners as much flexibility as possible while ensuring a “net conservation benefit” to the affected species. A net conservation benefit is defined in the policy as “the cumulative benefits of the management activities identified in a SHA that provide for an increase in a species’ population and/or the enhancement, restoration, or maintenance of covered species’ suitable habitat within the enrolled property, taking into account the length of the SHA and any off-setting adverse effects attributable to the incidental taking allowed by the ESP.”

The policy also states that the net conservation benefits “must be sufficient to contribute, either directly or indirectly, to the recovery of the covered species”, that “this contribution toward recovery will vary and may not be permanent”, and that a SHA “does not have to provide permanent conservation for (the) enrolled property”. With this in mind, it is clear that the policy’s intent is to provide landowners with a variety of options when determining what constitutes a net conservation benefit to a covered species. Furthermore, the FWS expects that the conservation benefits provided by any given landowner will vary, may contribute indirectly to recovery, and may not be permanent. Unrealized and short-term benefits are expected and are not excluded under the Safe Harbor policy. In fact, this system of permanent and short-term, and direct and indirect recovery benefits is what makes the Safe Harbor program and the SHAs developed under it so dynamic, popular with private landowners, and effective for listed species recovery.

Once the net conservation benefit determination is made, the SHA is finalized and the landowner applies for an ESP. Upon receipt of the SHA and ESP application, the FWS ensures that the SHA is in compliance with applicable laws and regulations and publishes a notice in the Federal Register announcing the availability of the SHA and soliciting comments on potential ESP issuance from the public. If the SHA is found in compliance with existing laws and regulations and there are no significant environmental effects that would result from implementation of the SHA, the FWS issues the landowner the permit. The landowner would then implement the specified conservation measures, annually report on the progress of the SHA, and notify the FWS if a reduction to baseline is necessary. The FWS is responsible for landowner compliance with the SHA and ESP and for monitoring the overall success of the Safe Harbor program.

A landowner can withdraw from the program at any time, but the regulatory assurances provided are valid only if the landowner is enrolled in and in compliance with the signed SHA. Once enrolled, private landowners are responsible for maintaining the habitat or individuals necessary to maintain the baseline and for conducting the conservation action(s) that will provide a net conservation benefit to the species. Often, this results in no significant changes to a given landowner’s management scheme. For example, many of the

41 participants in RCW Safe Harbor programs utilize long timber rotations and regular prescribed burns for bobwhite quail hunting, pine straw raking, and other activities. The use of fire and long timber rotations are practices that are beneficial to RCWs, so little more would be expected from these landowners in order to enroll in a Safe Harbor program, maintain their baselines, and provide a net conservation benefit. However, many have chosen to provide an increased conservation benefit to the species by installing artificial roosting/nesting cavities to encourage increases in their RCW populations. If new (i.e., above-baseline) groups of RCWs become established on the landowner’s property as a result of the enhancement activities, the landowner is not responsible for any additional management for these groups nor is the landowner liable for any incidental take of these additional RCW groups (since they would not be present except for the actions of the landowner). However, the landowner would need to notify the FWS is take of those above- baseline RCWs was anticipated.

RCW issues have not always been so easily remedied. In many cases, private forest land that was capable of serving as listed species habitat or could have been made into suitable habitat was not managed due to the perception that the presence of listed species would restrict traditional land uses or future development. As a result, there was a strong disincentive to listed species management such that landowners managed to reduce their risk of inhabitation by listed species. In the range of longleaf pine in the southeastern U.S., this included a prevalence of short rotation silviculture and elimination of natural fire regimes that was in certain instances intended to eliminate listed species such as the RCW or gopher tortoise (Gopherus polyphemus).

SAFE HARBOR SUCCESS STORIES IN LONGLEAF PINE As evidenced by the growth of RCW Safe Harbor programs, the availability of Safe Harbor has played a key role in removing the fear that has long surrounded listed species. Safe Harbor has encouraged increased management of longleaf pine stands that result in ecological characteristics favorable to the RCW and other species that inhabit these fire-maintained ecosystems. As of late 2000, regional Safe Harbor programs for RCWs have been implemented in the North Carolina Sandhills, east Texas, and southeastern Virginia, and state-wide programs have been permitted in South Carolina and Georgia. A number of individual landowner SHAs are also in development as are state-wide plans in Alabama, Florida, and North Carolina.

The existing programs in North Carolina, South Carolina, and Georgia have enrolled over 200,000 acres for 100 landowners, most of whom own longleaf pine stands on tracts from 3 acres to nearly 30,000 acres in size. This acreage also contains over 250 RCW family groups that are now subjected to increased levels of habitat management, including at least 15 new RCW family groups that were purposefully created as a result of these Safe Harbor programs. The level of protection for RCWs and longleaf pine stands is expected to increase in the short-term as new RCW SHAs are developed and the existing RCW Safe Harbor programs continue to grow.

CONCLUSIONS Safe Harbor programs are helping to create better working relationships between private landowners and those interested in longleaf pine and listed species conservation. Safe Harbor program enrollees are more often in close contact with federal and state biologists which creates lasting relationships between the public and private sectors relative to listed species management. A by-product of these relationships is that landowners have less anxiety when dealing with the involved agencies, which has helped alleviate negative feelings and fears about listed species and promoted increases in listed species conservation. Landowners have reacted positively toward the Safe Harbor and the regulatory and other incentives it provides. As a result, more acreage has been devoted to listed species-compatible management and an increasing number of landowners are receiving land management flexibility without the fear of additional regulatory burdens.

REFERENCES U.S. Fish and Wildlife Service. 1999a. Announcement of Final Safe Harbor Policy. Federal Register 64(116): 32716-32726. U.S. Fish and Wildlife Service. 1999b. Final Rule on Safe Harbor Agreements and Candidate Conservation Agreements with Assurances. Federal Register 64(116): 32706-32716.

42 RESTORATION OF LONGLEAF PINE HABITATS IN LOUISIANA AND EAST TEXAS BY THE NATURE CONSERVANCY

Richard Martin (The Nature Conservancy, P.O. Box 4125, Baton Rouge, Louisiana, 70821) Ike McWhorter (The Nature Conservancy, Silsbee, Texas)

INTRODUCTION An integral component of The Nature Conservancy’s (TNC) ecoregional conservation plans for the East Gulf Coastal Plain and Lower West Gulf Coastal Plain is restoration of degraded longleaf pine dominated habitats. Private landowners, public resource agencies and non-profit land management organizations have long tackled the challenge of restoring species composition and structure of longleaf systems. However, there are few published accounts of the thought process undertaken by longleaf restoration practitioners, nor is the network of people working on longleaf restoration sufficiently well-developed to ensure proper sharing of ideas and techniques.

The Louisiana and Texas offices of TNC have focused significant resources toward restoration of longleaf pine communities over the years and some of our experiences might prove useful to other land managers. Additionally, the sites currently being restored by TNC in Louisiana and Texas will add to the growing pool of experiments in longleaf restoration currently underway.

In addition to the authors, Wendy Ledbetter (Texas), Nelwyn McInnis, Susan Carr and Latimore Smith (Louisiana) are also deeply involved in longleaf restoration efforts being conducted by the Texas and Louisiana field offices of TNC.

Longleaf Sites Undergoing Restoration in Louisiana and Texas The Texas field office of TNC has been working on longleaf restoration since 1977 when Sandylands Preserve was established. At the current time, TNC of Texas is restoring longleaf habitats on four preserves (Table 1). Longleaf projects in Texas are in partnership with industrial timber companies and the majority of each site is dominated by upland longleaf communities.

Table 1. TNC longleaf pine restoration sites in Texas.

Site Size (ac) Ownership

Sandylands 5,685 TNC, Temple Inland Brushy Creek 2,000 International Paper Timberlake 1,540 TNC, Louisiana Pacific Little Rocky 930 TNC, Temple Inland

The Louisiana field office of TNC first entered the longleaf pine restoration arena with the acquisition of Lake Ramsay Preserve in 1994. TNC has acquired five longleaf-dominated preserves in Louisiana and has developed a close relationship with Louisiana National Guard by facilitating development and implementation of natural resources management plans (Table 2). In Louisiana, most of TNC’s longleaf restoration efforts have been focused on wet pine savannas and flatwoods, which are afforded little protection in comparison with upland longleaf habitats. In contrast to the strong partnerships TNC of Texas has forged with industrial timber companies, TNC of Louisiana has focused on working with private landowners and state agencies.

43 Table 2. TNC longleaf pine restoration sites in Louisiana.

Size Size (ac) Ownership

Money Hill Complex 4,396 TNC, private Camp Villere 1,500 LA National Guard Lake Ramsay 1,302 TNC, LA Wildlife and Fisheries. Camp Beauregard 974 LA National Guard Persimmon Gully 239 TNC, private CC Road Savannas 150 TNC

Establishing Restoration Objectives It is important that restoration practitioners consider all available data sources when formulating restoration objectives. The desired future conditions of any site should be based upon historic plant community composition, structure and distribution, priority conservation targets (species and communities) and the perceived likelihood of achieving success.

Historic or “pre-European settlement” conditions are often considered the end point for restoration projects. However, as many authors have pointed out “historic condition” is a moving target that is temporal in nature and it is probably more useful to understand the range of plant community conditions that were present at the site. Government Land Office records, which provide some insight into overstory composition, have been used by numerous restoration ecologists. The anecdotal comments within GLO records are often as informative as the witness tree data. Historic photos and site descriptions also provide insight about historic conditions.

While historic conditions of a site somewhat constrain the restoration limits for a site, of critical importance is a clear understanding of the species and plant community restoration targets. Careful examination of current composition and structure, especially distribution of long—lived indicator species (e.g., longleaf, pond cypress), and an understanding of the ecology and life history attributes of those species, are often indispensable sources of restoration guidance.

Historic conditions and conservation targets form the foundation for restoration plans but current conditions will dictate success. Significant resource allocation will be necessary to offset the loss of structuring forces such as hydrology, fire or gene flow and all restoration projects must be analyzed for costs and benefits. Conventional thinking dictates that some sites are beyond restoration (e.g., bedded plantation), However, there are few, if any, studies that have examined the restoration potential of highly altered sites. Although the Longleaf Alliance has provided an invaluable forum for exchange of ideas and experiences in the field of longleaf restoration ecology, the restoration literature is far from adequate.

TNC research at our Abita Creek Preserve in southeastern Louisiana is focusing on the restoration potential of lands that are approaching what many ecologists had once considered “critically degraded” and of suspect restoration potential. Because modern conservation requires a focus on restoration of functional landscapes, we are going to have to include restoration of heretofore “non restorable” sites in our suite of strategies. Initial results from two years of post-treatment (i.e., heavy logging) sampling at Abita Preserve have been exciting with notable increases in species richness, density of rare species, and total herbaceous cover, with little encroachment from non-native invasive species.

Those of us working to restore longleaf communities have focused a significant amount of effort toward identifying desired future conditions in a general sense; however, we have only recently began to establish specific benchmarks on which to measure our success. TNC of Louisiana has recently launched a protection and restoration effort focused on globally rare saline variant of wet longleaf savanna in southwestern Louisiana (Brimstone soils). Our specific restoration objectives were based upon a thorough review of the literature, consultation with longleaf ecologists, and our perception of “natural” conditions for this rare community type (Table 3).

44 Table 3. Restoration objectives for the saline variant of wet longleaf savanna in Southwest Louisiana.

Measure Objective Current Condition (Fall 2000)

Pine Overstory Cover 10-50% (heterogeneous) 6% Longleaf Composition 70-90% (skewed toward 90%) 16% Hardwood Overstory Cover 1-5% 10% Hardwood Species Richness > 4 species per 100m2 5 species per 100m2 Hardwood Understory Cover 5-15% 19% Herbaceous Cover 90-100% 74% Proportion Graminoids 50-90% 68% Proportion Forbs 10-50% 7% Total Native Species Richness > 10 species per m2 11 species per m2

Although most experienced longleaf managers could develop an acceptable restoration plan for this site, clearly identifiable and biologically defensible restoration targets will allow for true adaptive management and should result in efficient allocation of scarce management dollars.

CONCLUSIONS Development and implementation of successful restoration plans require practitioners to consult all available sources of data. In addition to formulating generic desired future conditions, it is important to establish objective, defensible restoration targets on which to measure success and to serve as the foundation for adaptive management. However, in spite of our efforts, restoration ecology remains as much an art as a science and we will continue to rely on our best guesses until the volume of published studies of restoration projects across the range of longleaf increases significantly.

45 LONGLEAF PINE SEEDLING PRODUCTION

James P. Barnett (USDA Forest Service, Southern Research Station, 2500 Shreveport Highway, Pineville, LA 71360)

ABSTRACT: Longleaf pine is a highly desirable species, resisting fire, insects and pathogens, and produces quality solid-wood products, but regeneration of the species has been difficult. Natural regeneration is feasible only on a small portion of the area considered to be longleaf pine type. Therefore, artificial regeneration must become a reliable means of regenerating the species if restoration of the ecosystem is to be successful. The knowledge and technology to reestablish longleaf pine by planting bareroot stock has improved significantly within the last decade. However, numerous studies show that container seedlings survive and grow better than bareroot stock. Additional studies are underway to refine these techniques and provide a better understanding of the unique physiological attributes of longleaf seedlings that may allow us to improve regeneration success by planting.

INTRODUCTION Longleaf pine was widely distributed in the presettlement forests of the southern Coastal Plain, but now occupies less than 5 percent of its original range. It is a highly desirable species, resisting fire, insects and pathogens, and produces quality solid-wood products. Because of its limited distribution, natural regeneration is feasible only on a small portion of the area considered to be longleaf pine type. Artificial regeneration by direct seeding or planting has been difficult, but renewed interest in longleaf pine has caused us to reevaluate these approaches to stand establishment. The difficulty in regenerating longleaf is related to its unique botanical characteristics: (1) low and infrequent seed production, (2) a seedling “grass” stage characterized by delayed stem elongation, (3) poor storability of bare-root nursery stock that results in low survival, and (4) seedling intolerance to shade conditions caused by competition (Barnett and McGilvray 1997).

Direct seeding is an inexpensive reforestation option, particularly suitable for small landowners, because spot or broadcast seedling by hand equipment is easy and can be effective (Derr and Mann 1971). However, the use of direct seeding has declined because it is less reliable than planting, particularly of container stock, and because the seed coat repellent for rodents is not longer available. Thiram is an effective bird repellent, but endrin, which is registered for rodent protection is no longer manufactured in the United States. Recent studies indicate that oleoresin capsicum, in combination with thiram, effectively protects longleaf pine seeds from bird and rodent predation (Barnett 1998, Nolte and Barnett 2000). Broadcast seeding is probably not feasible due to the cost of longleaf seeds, but spot seeding is an acceptable option for small landowners.

The key to successful artificial regeneration is availability of quality seeds. Obtaining adequate quantities of good quality longleaf pine seeds requires extraordinary efforts.

SEED QUALITY AND QUANTITY ISSUES Use high-quality seed (viability >80 percent is desired) to obtain uniformity in bareroot nursery beds and to reduce costs of sowing multiple seeds and thinning to one seedling per cavity in container nurseries. Obtaining high-quality seeds remains a problem. To achieve good seed quality, organizations inolved in the collection and processing of longleaf seeds need to pay particular attention to cone maturity and collection and seed processing techniques (Barnett 1997). Longleaf seeds commonly have significant populations of pathogenic fungi that result in seedling mortality, so a fungicidal seed treatment is generally effective in reducing early seedling mortality. A ten-minute soak in a benomyl solution (2 tbsp/gal) reduces fungal infestations and improves germination and establishment (Barnett et al. 1999).

BAREROOT SEEDLING PRODUCTION Longleaf pine seedlings have no early epicotyl growth and are, therefore, very sensitive to competition. The seedlings’ initial height growth occurs most quickly in full sunlight. Studies have shown that site preparation that delays competition development over an extended period—prescribed fire 1- or 2-years after planting and/or post-planting competition control by mechanical or chemical means—will allow height initiation within 1 to 3 years after planting.

46 The knowledge and technology to reestablish longleaf pine by planting bareroot nursery stock have improved significantly in the last decade. The components of successful regeneration include: (1) well-prepared, competition-free sites, (2) healthy, top-quality, fresh planting stock, (3) meticulous care of stock from lifting to planting, (4) precision planting, and (5) proper post-planting care (Barnett and Dennington 1992). It is essential that all five of these elements come together for successful planting of bareroot stock. It is difficult to control all of these factors, therefore, planting success with bareroot longleaf pine stock remains elusive. Although these components for successful regeneration apply to container stock, success in establishment is markedly better with container material (Boyer 1987).

CONTAINER SEEDLING PRODUCTION Numerous studies have demonstrated that under adverse planting conditions, such as, poor sites, conditions of moisture stress, and out-of-season planting, container seedlings survive and grow better than bareroot stock. This is generally attributed to the fact that the root systems of container plants remain intact, with those of bareroot plants are severely damaged during lifting. Thus, container stock has a significantly shorter period of transplant shock or readjustment than bareroot seedlings.

Longleaf pine seedlings can be grown in the open without a structure, but if crops are overwintered, polyethylene or other protective covering may be needed to protect seedlings from strong desiccating winds and temperatures below 25°F. An adequate watering system is essential for container production—a simple, stake type is with sprinkler heads is normally sufficient. The ideal container cavity should have a volume of about 6 cubic inches, a minimum depth of 4 inches, and a seedling density of <50 per square foot. A growing mix of sphagnum peat, #2 grade horticultural vermiculite, and possibly a small percentage of perlite has been a consistently good product for filling the containers. The pH of the medium should be adjusted, if necessary, to about 4.5 to 5.0 and most growers incorporate a slow-release fertilizer to reduce the frequency of fertilizer applications during the growing phases (Barnett and McGilvray 1997).

The best growing schedule, both biologically and economically, for longleaf pine is to sow seeds in the spring, grow through the summer, harden the seedlings naturally in the fall, and outplant them in late fall or early winter. The best strategy is to sow one seed per cavity, but excellent seed quality is essential for this option. If viability is below about 80 percent, two seeds should be sown per cavity and then if two germinate, one should be thinned about the time the seedcoats shed. Water management is a critical aspect of seedling culture and experience is the best way to learn proper watering techniques. More details of cultural practices are found in Barnett and McGilvray (1997).

Despite their bulk and weight, container seedlings are easy to plant by hand or machine because their root systems are uniformly shaped. The control of planting depth is critical for longleaf pine. The bud should be at about the soil surface. Dibbles shaped like the root plug work well because the problem of planting too deep can be avoided.

CONCLUSIONS Regeneration by direct seeding or planting has not been highly successful, so the renewed interest in longleaf pine has caused us to reevaluate these approaches to seedling establishment. The knowledge and technology to reestablish longleaf pine by planting bareroot stock has improved significantly in the last decade. However, establishment of these seedlings remains more difficult than for container stock.

Reforestation success can be improved significantly by planting seedlings grown in containers. Container stock survives better than bareroot stock on typical longleaf pine sites and the length of time seedlings stay in the grass stage is reduced. However, using container stock does not eliminate the critical need for controlling competition during the first growing season. Such competition control ensures that seedlings begin height growth during the second year following planting.

LITERATURE CITED Barnett, J.P. 1997. Improving the quality of longleaf pine seed from orchards. In: Proceedings, 24th biennial southern forest tree improvement conference; 1997 June 9-12; Orlando, FL. Southern Forest Tree Improvement Sponsored Pub. 46. Gainesville, FL: University of Florida, School of Forest Resources and Conservation: 124-132.

47 Barnett, J.P. 1998. Oleoresin capsicum has potential as a rodent repellent in direct seeding longleaf pine. In: Waldrop, T., compl. Proceedings ninth biennial southern silvicultural research conference; 1997 February 25-27; Clemson, SC. Gen. Tech. Rep. SRS-20. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station: 326-328. Barnett, J.P.; Dennington, R.W. 1992. Return to longleaf. Forest Farmer 52: 11-12. Barnett, J.P.; McGilvray, J.M. 1997. Practical guidelines for producing longleaf pine seedlings in containers. Gen. Tech. Rep. SRS-14. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 28 p. Barnett, J.; Pickens, B.; Karrfalt, R. 1999. Improving longleaf pine seedling establishment in the nursery by reducing seedcoat microorganisms. In: Haywood, J.D., compl. Proceedings of the 10th biennial southern silvicultural research conference; 1999 February 16-18; Shreveport, LA. Gen. Tech. Rep. SRS-30. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station: 339-343. Boyer, William D. 1989. Response of planted longleaf pine bare-root and container stock to site preparation and release: fifth-year results. In: Miller, J.H., compl. Proceedings of the 5th biennial southern silvicultural research conference. Gen. Tech. Rep. SO-74. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station: 165-168. Derr, H.J.; Mann, W.F., Jr. 1971. Direct seeding pines in the south. Agricultural Handbook 391. Washington, DC: U.S. Department of Agriculture, Forest Service. 68 p. Nolte, D.L.; Barnett, J.P. 2000. A repellent to reduce mice predation of longleaf pine seed. International Biodeterioration and Biodegradation 45(3/4): 169-174.

48 PLANTING METHODS: IS BROWN DOWN, GREEN UP GOOD ENOUGH?

Dale C. Pancake, Jr. (Solon Dixon Forestry Education Center)

For many decades longleaf pine has had the reputation of being difficult to regenerate. Problems with seedling survival and delayed height growth have caused people to plant other species – even where longleaf may have been best suited for the site. Recent gains in seedling quality and a better understanding of the requirements of the species greatly improve the chances of success. Successful artificial regeneration of longleaf depends on careful attention to the following planting premises.

1. Plant good planting stock: For containerized seedlings, that means a seedling with at least a 3.5 inch plug, a ¼ inch root collar diameter and a firm plug held together by many roots. There should be no circling roots visible around the outside of the plug. The needles should be at least 4 inches long and have good, healthy green color. Bare root seedlings should have a root collar diameter of at least one half inch, numerous lateral roots and healthy needles. Best success is achieved when bare root seedlings are planted fresh, that is, within a week of lifting in the nursery. Survival has been shown to drop off with longer storage time. Bare root longleaf seedling do best when machine planted because of the difficulty in properly hand planting the large root system of these seedlings. Typically bundles of run- of-the-bed bare root seedlings contain a variety of size classes. If the seedlings are not graded at the nursery, then planters would do well to discard the very small seedlings (<1/4 in.) in the field. Very large seedlings of ¾ inch RCD or larger require special care to ensure that they are properly planted.

2. Use an experienced planting crew. This means planters that have successfully planted longleaf in the past and that understand that it is critically important that the seedlings be placed properly in the ground.

3. Plant in well-prepared ground. Good site preparation is a must. Weed competition, especially grass competition should be dealt with before planting. Longleaf seedlings are very intolerant of root competition and overtopping weeds or grasses. On cutover areas, site preparation can be accomplished mechanically, by burning, by chemicals, or by combinations of these treatments. Any physical obstacles to good planting access should be removed as well. On agriculture fields with a plow-pan or hardpan, ripping or sub-soiling is recommended to allow good root penetration and moisture movement through the soil. Prior to planting in pastures or following peanuts or other legume crops where white-fringed beetle larvae are present, scalping is recommended. This treatment peels back a strip of sod or surface soil 3-4 inches thick and provides a competition-free, insect-free strip three feet wide for planting. After ripping, the soil should be allowed to settle before planting. Usually 3-4 good soaking rains are needed to adequately settle the rips. Seedlings can be planted slightly offset from the rips to avoid future collapse or settling. The seedling roots will find the rip as they follow moisture in the soil.

4. Proper handling of the seedlings. Poor handling practices can lead to unnecessary mortality or even planting failure. Seedling bundles or containers and trays should be protected from sun, wind, dry air and freezing. Roots should be protected at all times.

5. Plant in moist soil. Adequate soil moisture both at the surface and in the subsoil is needed at the time of planting for good survival. Dry soil equals dry seedlings equals dead seedlings!

6. Proper planting: Containerized longleaf pine seedlings can be planted either by machine or by hand. Bare root seedlings should only be machine planted. In machine planting , the seedlings are either placed directly in the soil by the person riding the planter or they are placed in fingers on the machine and the machine places them in the soil. In either case, the operator is the key to a successful planting operation. If the seedlings are properly planted at the right depth, chances of success are good. Poorly planted seedlings greatly diminish chances of survival. A properly machine planted seedling has the roots evenly distributed in the planting slit and the bud right at the soil surface. If the bud is buried, it cannot open to produce new needles. Deeply planted seedlings are doomed to decline and die. Seedlings planted too shallowly, with the root collar and part of the tap root exposed, also do poorly since they are exposed to sun, wind, drying and freezing. When starting a machine planting operation, it is important to plant a trial row or two to make sure the machine is properly set for the soil conditions on

49 the site and that the seedlings are being planted at the proper depth. Hand planting of containerized seedlings can be accomplished using dibbles, plug-type tools or even planting shovels. The size of the hole opened by these tools should match the size of the plug on the seedling so that the plug is straight and the bud is left right at the surface of the soil.

Planting myths a) “Long roots on bare root seedling should be pruned to facilitate planting.” False! This practice may make it easier to plant the seedlings but it is often fatal to the seedling! The seedlings have been shocked already by lifting from the nursery. If we then cut off their roots, we remove that portion of the seedling that is needed to begin absorbing moisture after planting. b) “Plant them deep, the soil will settle.” False! Other pines that have a stem can be planted at varying depths without affecting survival. Longleaf seedlings lack a stem. The bud is right on the root collar and must be planted right at the soil surface to live. c) “Long laterals should be wrapped around the roots in order to get all of the roots in the planting hole”. False! This will set up girdling roots from the start. Roots should be planted straight in the hole. d) “Trail the roots in the slit when machine planting.” This can result in a tree with a uni-directional root system in a single plane that will not be capable of supporting itself. “Tumping over” of longleaf saplings often can be traced back to this.

Planting depth In a study conducted by the Longleaf Alliance, containerized seedlings were planted at four depths: a) with the plug 1 cm above the soil surface, b) with the plug exposed at the soil surface, c) with the plug buried 1cm below the surface but with the bud right at the surface, and d) with the plug buried 2 cm below the soil surface with the bud buried. At the end of the first season nearly 40 percent of the buried seedlings had died compared with less than 15 percent for the other treatments. At the end of the second season, between 30 and 48 percent of the seedlings with above surface buds had started height growth compared with zero seedlings initiating height growth for those with the bud buried 2 centimeters.

CONCLUSION We spend a great deal of money and effort on site preparation, seedlings, planting, and release. Proper Planting Depth Is Critical For Success.

50 COST SHARE AND INCENTIVE PROGRAMS

Don Feduccia (Louisiana Department of Agriculture and Forestry)

What is FPP? The Forestry Productivity Program (FPP) provides financial assistance to eligible landowners for establishing and improving a crop of trees. This program helps offset a landowner's expenses by sharing the cost of implementing specific forestry practices designed to produce a timber crop. Cost-share payments cover 50 percent of the total cost of implementing one or more forestry practices, not to exceed a maximum limit set for each individual practice. Eligible landowners can receive up to $10,000 of FPP assistance each fiscal year.

Who Is Eligible for Assistance? FPP assistance is available to landowners in all Louisiana parishes who own a minimum of five (5) contiguous acres suitable for growing a commercially valuable timber species. There is no maximum size of ownership limiting a landowner's participation in the program. Applications for FPP assistance are accepted on a first-come, first-served basis.

Ten-year Requirement Each landowner who participates in a FPP cooperative agreement with the Louisiana Department of Agriculture and Forestry (LDAF) shall be required to maintain the forestry usage for a period of 10 years. If land is sold, conveyed, etc. before the end of the agreement, the original recipient of funds is bound unless the new owner assumes responsibility in writing.

Six Steps to FPP 1. Application Applications are available from LDAF foresters, industry foresters, forestry consultants or foresters from other governmental agencies. A non-refundable application fee is required.

2. Management Plan A professional forester trained in LA's FPP can prepare a management plan stating practice(s) and acres requested.

3. Approval Plan and requested practice(s) are approved by LDAF Forester and application forwarded to State Office.

4. Authorization Landowner will receive a notification letter that practice(s) funding has been authorized. Landowner is responsible for making the necessary arrangements to implement each practice (e.g., contract with vendor, order seedlings, etc.). Landowner MUST notify LDAF Forester prior to beginning any practice.

5. Completion As each practice is completed, the landowner notifies the LDAF Forester, who then makes final inspection of the work to ensure that the practice(s) has been applied in compliance with FPP standards.

6. FPP Payment Landowner provides LDAF Forester with copies of all receipts pertaining to practice(s) completed. This information is forwarded to State Office for payment.

51 Forestry Practices Eligible for FPP Assistance

Planting or Seeding

Objective: To establish a crop of trees by planting or seeding pine or hardwood. This practice includes the cost of seedlings, seed, labor, and if needed, site preparation.

Post-planting site preparation is allowed to reduce or control undesirable vegetation within the first growing season of an established crop of trees.

Site Preparation for Natural Regeneration

Objective: To prepare an area in order for natural regeneration to become established. This practice includes the appropriate method(s) of site preparation.

Control of Competing Vegetation

Objective: To release an existing crop of desirable trees from undesirable vegetation. This practice includes the application of chemical, pre-commercial thinning, and release of longleaf pine seedlings by burning.

52 A BRIEF OVERVIEW OF FIRE AND SEASON OF BURN IN NATIVE LONGLEAF PINE ECOSYSTEMS

Sharon M. Hermann (Research Associate, Auburn University, Auburn, AL 36849)

For much of the first half of the last century, fire in longleaf pine (Pinus palustris) forests was viewed as detrimental to both trees and wildlife. However more recently, observations and scientific research have demonstrated that fire, when applied properly, is actually beneficial to longleaf trees. In fact, we now understand that fire suppression has been perhaps the single most important factors in the decline of a biome that once dominated much of the uplands in the southeastern coastal plain (Frost 1993, Simberloff 1993). Unfortunately there is less awareness about the significance and use of different times of burning in this native ecosystem, especially of fires in the growing- or lightning-season.

An important consideration for management of longleaf pine forests is not if fire should be used, but rather when it should be applied. This question should be addressed in the context of long-term planning of a burn regime. In its simplest form, a burn regime consists of two components: frequency and seasonality. This paper focuses on seasonality. Frequency is often a region- and/or site-specific issue. In reality, there are many other factors to be considered in a long-term fire plan, including day-of-burn conditions (weather, time of day, etc.) and ignition pattern. However, season of burn is one of the most important components of prescribed fire, and until recently, was often not considered. There has been a mistaken belief that altering season of burn was an issue only for nature preserves. Although parks were some of the first agencies to expand their burning programs (Stevenson 1998), forests managed for timber and/or wildlife can also profit from alternative applications of fire. Prescribed lightning-season burns have been introduced on national forests (Ferguson 1998) and national wildlife refuges (Ingram and Robinson 1998) with impressive success. On-going research and observations of land managers indicate that expanding the dates used for prescribed fire can have many advantages for longleaf pine forests. The basis for these positive effects resides in the ecology of the longleaf pine ecosystem.

Native longleaf pine forests are defined by the single dominant tree species growing in what, for hundreds of years, has been described as an open park-like forest. The trees are usually represented by multi-age (-size) classes and clustered in patches under an open-canopy (Platt et al. 1988a). Vegetation complexity of this forest type resides in the spatial arrangement of the trees and in the hundreds of plant species in the native ground cover. When the ground cover is in a relatively natural state, it is dominated by grass and forb species, the vast majority of them perennials. The grasses are of particular significance for fire management because these species, plus pine needles, constitute the fine fuel that is critical for effective burning. The term longleaf pine forest actually covers many variations of similar vegetation types. Harcombe et al. (1993) provide a review of vegetation in the West Gulf Coastal Plain while Peet and Allard (1993) cover the Southern Atlantic and Eastern Gulf Coast Regions. Although these regions generally support a variety of similar species of ground cover plants, the three-awned wiregrass, Aristida stricta (including A. beyrichiana of Peet 1993) is absent in areas west of the Mississippi River. This species is significant partly because of its often-discussed flowering response to lightning-season fire (see below).

Historically, prescribed fire was set during the winter or early spring months (usually ending in March or early April), normally a few days after rain and the passage of a weather front. Justifications for setting fire during this time period are listed in Table 1. On areas used for quail hunting burning was often limited to late February to late March or early April. This short time period comes after the end of quail hunting but before the birds begins to nest. Although quail readily re-nest, land managers have been concerned that populations would suffer if some early nests were lost. All of the reasons for burning in the dormant season are based, in part, on reality. However these assumptions do not negate possibilities for applying lightning- season fires (Table 2).

In the last twenty years, there has been increased understanding of the significance of fire at different times of the year for many components of the native forests, including effects on longleaf pine, wildlife, insects, ground cover plant species, and hardwood control (Table 3). The publication “Seasonal effects of prescribed burning in Florida: a review” (Robbins and Myers 1992) provides a very readable, general overview of available literature on fire effects in a variety of southern pine forests. A review paper by Streng et al. (1993)

53 summarized available scientific literature on season of burn research specifically related to longleaf forests, critiqued the methods used in each study and offered additional data on ground cover plant responses. Hermann et al. (1998) also discussed plant flowering and changes in insect abundance related to month of burn. Hiers et al. (2000) focused on legume responses and Platt et al. (1988b) examined the time of flowering of the total ground cover plant community.

Both Robbins and Myers (1992) and Streng et al. (1993) stressed the importance of specifying month when discussing issues related to date of burn. Terms such as “warm and cool season”, “summer and winter”, or “growing and dormant season” are vague and may be interpreted as different months especially when the entire range of longleaf (southern Virginia to eastern Texas to central Florida) is considered. Lightning- season is a specialized term defined by local weather patterns that correspond to time periods when lightning- ignited burns are most likely to occur. Although, lightning can start fires during any month, their likelihood varies with weather, geography, season, and among years (Komarek 1964, Olson and Platt 1995). As a general rule, they are concentrated sometime between mid-April through July or August, depending on rainfall. July is one of the wettest months across much of the range of longleaf pine, although severe droughts can occur at that time.

LONGLEAF PINE An important consideration for any property owner with longleaf pine is whether lightning-season burns will harm trees, especially recruits. Research results from many studies indicate that when the habitat is in good condition (see below) and the trees are healthy, there is little or no adverse effect on most sizes of longleaf pine trees (Robbins and Myers 1992, Streng et al. 1993, and Glitzenstein et al. 1995). Juvenile trees may actually benefit from fire at this time of the year. For example, under a frequent fire regime, Grelen (1983) documented that five-year old grass-stage longleaf pine grew better and had lower mortality when burned in May compared to other fire treatments. In addition, most individuals bolting out of the grass-stage were not adversely affected (Grelen 1983), although many foresters still caution against burning when many small trees are in this phase. Streng et al. (1993) and Glitzenstein et al. (1995) documented no over all difference in mortality or growth of longleaf pine associated with differences in season of burn. However, they did discover that fall (October and November) fires kill more, small longleaf on sandhill sites and winter (January and February) fires kill more individuals in small size classes on flatwoods sites (Streng et al. 1993, Glitzenstein et al. 1995).

There are come cautionary points to consider. Before adding lightning-season burns to a fire regime, the fuel condition should be carefully evaluated, especially at the base of large trees. Over time, cool winter burns, coupled with elevated soil and litter moisture at the base of large trees, may result in a buildup of pine needles near the trunk. If not removed by raking or other means, the first prescribed lightning-season fire may kill the cambium, a tree’s live tissue just under the back. Additionally, one- and often two-year old longleaf seedlings do not survive fire in any month.

FOREST STRUCTURE AND HARDWOOD CONTROL A vegetation structure defined by open sunlit areas is necessary for most, if not all, of the plant and animal species endemic to longleaf pine forests. Burning prunes the woody stems of ground cover oaks; these species are fire-adapted, producing acorns on knee-height stems. It also depresses the encroachment of hardwood species from nearby, non-fire ecosystems such as drains and hardwood forests. Fire is a cost effective means for maintaining the appropriate longleaf pine habitat structure and lightning-season burns can contribute to the financial benefits (Ferguson 1998). In north Florida sites with intact longleaf pine-wiregrass forest, Streng et al. (1993) demonstrated that annual or biennial fires in April and May over a eight year period produced over 80% topkill of various oak species stems of all sizes, compared to less than 50% for other seasons of burns. In addition, fires applied in the same frequencies and seasons resulted in the highest likelihood of completely death of hardwood stems compared to burns in other months. For early lightning- season fires, total mortality of oak species ranged from almost 30% for small stems (0.75-2.0 in DBH) to approximately 80% for larger stems (> 4 in DBH).

BOB WHITE QUAIL AND WILDLIFE FOOD RESOURCES Many land owners and mangers are concerned about the possibility of lightning-season fires harming wildlife species. Despite widespread belief that ground-nesting birds will suffer if fire occurs during nesting periods,

54 this may not be true. Although some nests may burn, no decrease in bob white quail populations was observed over three years at sites in North Florida burned during February and May (reported in Hermann et al. 1998). In addition, both plant and insect resources eaten by quail remained the same or increased after burning in May or June (summer) compared to January or February (winter). Although insects were in low abundance at the beginning of the summer (June) on areas burned just a few weeks earlier compared to winter fires, by July or August on summer burned sites greatly exceeded winter burned areas (Hermann et al. 1998).

NATIVE GROUND COVER PLANTS As noted earlier, native ground cover in longleaf pine forests contains hundreds of plants. Not surprisingly these species do not all react in the same manner to timing of burn. Although Platt et al. (1988b) demonstrated an overall shift in timing in flowering of the ground cover plant community related to month of burn, research on specific plant species has revealed a range of responses. Flower and/or fruit production of at least a few species has been shown to be elevated for each month of burn that has been studied (Hermann et al. 1993, Hiers et al. 2000, Streng et al. 1993). It is obvious that effects of season of burn on ground cover species are complex, so it is instructive to consider results for three-awned wiregrass, one of the most common species in the longleaf pine ground cover east of the Mississippi River. This species may be one of the most fire-adapted grasses in North America. It normally produces almost no flower or seed unless burned in the lightning-season (Streng et al. 1993). Other grass species display similar responses, including Florida dropseed (Sporobolus floridanus), however some species show little response to different months of burn (Streng et al. 1993).

It might seem as if the information on ground cover plant species and season of burn presents a confusing, perhaps even contradictory picture. Although much research remains to be done, at least three unifying factors emerge from current information: 1) the vast majority of ground cover plants are perennials so extensive seed production is not necessary every year, 2) wiregrass, the dominant ground cover species in a large portion of the range of longleaf pine, responds positively to lightning-season fire, and 3) lightning- season burns appear to be most effective in controlling hardwood stems. This final point may be the most significant one. A low-growing ground cover is a critical part of the vegetation structure required to maintain the rich biotic diversity of longleaf pine forests. Restoration and management of the ground cover is one of the biggest challenges for effective fire management. While species of special concern should always be monitored under any fire regime, inclusion of lightning season burns in a prescribed fire regime should enhance most long-term management goals in this native forest.

ADDITIONAL BENEFITS OF BURNING IN THE LIGHTINING-SEASON Some benefits of prescribing lightning-season burns are not easy to quantify. These fires can burn farther into wet areas that adjoin upland longleaf pine forests compared to dormant season burns. This can create a more natural boundary between these two ecosystems and produce better conditions for rare plants that live in the ecotone (Ingram and Robinson 1998). An additional simple but possibly significant advantage of adding lightning-season fires to a land management plan is that is that more days become available for burning every year. In addition, burns during the lightning-season may provide alternative conditions for smoke management. Prevailing wind patterns often differ among seasons of the year. Areas, such as roads, that may be readily inundated by smoke one season may be easier to manage under different weather conditions in another season.

SOME CONSIDERATIONS FOR DETERMINING OF LIGHTINING-SEASON FIRES ARE APPROPRIATE FOR A SITE Lightning-season fires have the potential to benefit almost every type of longleaf pine site but careful evaluation of current conditions and a clear understanding of long-term management goals are essential for development of a successful burn program. As noted above, inclusion of lightning-season fires in a burn program can enhance many land management goals but it may not compensate for prolonged periods of fire suppression and/or other types of human activities. Some of the necessary points to consider when altering a prescribed fire regime are summarized in Table 4.

Although fires in the lightning-season were a natural part of longleaf pine forests before humans altered the ecosystem, applying such a fire in a modern landscape without considering the site’ current condition could

55 have undesirable consequences. Vegetation structure, plant species composition and/or fuel loads may have changed due to past fire suppression or long periods of only dormant season burns. Altered fuel load is often a problem. As noted above, cool or dormant season fires may result in an accumulation of litter at the base of large trees. Excess must be eliminated before introducing lightning-season. This can be accomplished by raking around the trees or, in some cases, through a series of dormant-season and early spring burns designed to target litter removal. Vegetation structure is another factor to consider. Past logging and/or fire suppression may have resulted in many small woody stems and/or patches devoid of fine fuel. Even aggressive application of fire may not rectify all management problems in a reasonable time frame. One time applications of additional techniques may be useful in “jump-starting” fire management. Selective cutting, girdling, herbicide use and planting may be useful habitat restoration tools but care should be exercised so as not to further damage remnant native vegetation.

The presence of exotic species must be considered in any site management plan. Although fire, especially in the lightning-season, may suppress many exotic species there are others that are fire adapted. Cogongrass (Imperata cylindrica) is an invasive exotic species that thrives under lightning-season fire management, can alter the effects of burns and produces alleopathic compounds that may create monotypic stands. Once established, herbicide may be the only way to eradicate such plants.

In addition to evaluating potential changes in vegetation and fuels before re-introducing lightning-season burns, possible differences in fire weather must be considered. In the Southeast, predictable, low-velocity winds are most likely to occur during the dormant season (fall through early spring). However, differences in wind patterns and resulting shifts in fire and smoke behavior are not inherently bad unless they are unexpected. It’s best to avoid all surprises. If the general public is likely to be concerned about a site, an education program, prior to the re-introduction of lightning-season fire, will be well worth the effort.

CONCLUDING THOUGHTS If there are concerns about expanding the months over application of fire (fuel load, fire severity, weather patterns, etc.), the shift may be facilitated by first implementing a series of dormant season burns. The first sites targeted for introduction of lightning-season fires should be small to moderate size and in relatively good ecological condition. As burn planners and fire crews become more experienced, the size and complexity of burn blocks can be increased. With careful planning and prudent application, lightning-season fires will help maintain high quality native longleaf pine forests at their current condition and aid in the restoration of degraded areas. Although almost any month of burn will produce a positive response in at least a few species, lightning-season burns are the most effective in controlling hardwoods.

In addition to immediate concerns of how to expand the season of burn on a site, there are issues related to long-term plans. Fire return interval (frequency) is the most common component of a prescribed fire regime outlined in a long-term plan. By projecting both frequency and months of burn, prescribed managers can develop priorities for application of a broader range of types of fires for each burn unit, ranging from periodic to almost exclusive use of lightning-season fire. Planning enhances management goals as well as efficient use of fire resources, including the number of burn days available each year.

REFERENCES Ferguson, J.P. 1998. Prescribed fire on the Apalachicola Ranger District: the shift from dormant season to growing season and effects on wildfire suppression. pp. 120-126 in T.L. Pruden and L.A. Brennan (eds.) Proceedings Tall Timbers Fire Ecology Conference No. 20: Fire in ecosystem management: shifting the paradigm from suppression to prescription. Frost, C.C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. pp. 17-43 in S.M. Hermann (ed.) Proceedings Tall Timbers Fire Ecology Conference No.18: The longleaf pine ecosystem: ecology, restoration and management. Glitzenstein, J.S., W.J. Platt, and D.R. Streng. 1995. Effects of fire regime and habitat on tree dynamics in North Florida longleaf pine savannas. Ecological Monographs 65:441-476. Harcombe, P.A., J.S. Glitzenstein, R.G. Knox, S.L.Orzell, and E.L. Bridges. 1993. Vegetation of the longleaf pine region of the West Gulf Coastal Plain. pp. 83-104 in S.M. Hermann (ed.) Proceedings Tall Timbers Fire Ecology Conference No.18: The longleaf pine ecosystem: ecology, restoration and management.

56 Hermann, S.M., T. Van Hook, R.W. Flowers, L.A. Brennan, J.S. Glitzenstein, D.R. Streng, J.L. Walker, and R.L. Myers. 1998. Fire and biodiversity: studies of vegetation and arthropods. Transactions of the North American Wildlife and Natural Resources Conference 63:384-401. Hiers, J.K., R. Wyatt and R.J. Mitchell. 2000. The effects of fire regime on legume reproduction in longleaf pine savannas: is season selective? Oecologia 125:521-530. Grelen, H.E. 1983. May burning favors survival and early height growth of longleaf pine seedlings. Southern Journal of Applied Forestry 7:16-19. Ingram, R.P. and D.H. Robinson. 1998. Evolution of a burning program on Carolina Sandhills National Wildlife Refuge. pp. 161-166 in T.L. Pruden and L.A. Brennan (eds.) Proceedings Tall Timbers Fire Ecology Conference No. 20: Fire in ecosystem management: shifting the paradigm from suppression to prescription. Komarek, E.V. 1964. The natural history of lightning. pp. 139-189 in E.V. Komarek (ed.) Proceedings Tall Timbers Fire Ecology Conference No. 3. Olson, M.S. and W.J. Platt. 1995. Effects of habitat and growing season fires on resprouting of shrubs in longleaf pine savannas. Vegetation 119:101-118. Peet, R.K. 1993. A taxonomic study of Aristida stricta and Aristida beyrichium. Rhodora 95:25-37. Peet, R.K. and D.J. Allard. 1993. Longleaf pine vegetation of the Southern Atlantic and Eastern Gulf coast regions: a preliminary classification. pp. 45-81 in S.M. Hermann (ed.) Proceedings Tall Timbers Fire Ecology Conference No.18: The longleaf pine ecosystem: ecology, restoration and management Platt, W.J., G.W. Evans, and S.L. Rathbun. 1988a. The population dynamics of a long-lived conifer (Pinus palustris). American Naturalist 131491-525. Platt, W.J., G.W. Evans, and M.M. Davis. 1988b. Effects of fire season on flowering of forbs and shrubs in longleaf pine forests. Oecologia 76:353-363. Robbins, L.E. and R.L. Myers. 1992. Seasonal effects of prescribed burning in Florida: a review. Tall Timbers Research, Inc. Miscellaneous Publication No. 8. Simberloff, D. 1993. Species-area fragmentation effects on old-growth forests: prospects for longleaf pine communities. pp. 1-13 in S.M. Hermann (ed.) Proceedings Tall Timbers Fire Ecology Conference No.18: The longleaf pine ecosystem: ecology, restoration and management. Stevenson, J.A. 1998. Evolution of fire management in Florida's state parks. pp. 99-101 in T.L. Pruden and L.A. Brennan (eds.) Proceedings Tall Timbers Fire Ecology Conference No. 20: Fire in ecosystem management: shifting the paradigm from suppression to prescription. Streng, D.R., J.S. Glitzenstein and W.J. Platt. 1993. Evaluating effects of season of burn in longleaf pine forests: a critical literature review and some results from an ongoing long-term study. pp. 227-263 in S.M. Hermann (ed.) Proceedings Tall Timbers Fire Ecology Conference No.18: The longleaf pine ecosystem: ecology, restoration and management.

57 Table 1. Historical justifications for limiting the application of fire in longleaf pine forests to dormant season months (November through March).

1) Predictable wind patterns follow seasonal cold fronts. 2) Brown (dead) and/or lack of green (live) vegetation adds to the fine fuel base. 3) Elevated soil moisture can create a patchy burn pattern. 4) “Dormancy” (quiescence) of many plants and , including ground-nesting birds may minimize injury, mortality and/or loss of active nests. 5) Lower air temperature makes work conditions more comfortable for employees. 6) Overall weather is thought to produce less severe effects from fire. ______

Table 2. Comparisons between lightning-season fires (often April to July or August) and the historical justifications for use of dormant season burns cited in Table 1.

1) Winds are often less predictable in the lightning-season but are not always problematic. 2) Disturbed or old-field vegetation may burn best in the dormant season, depending on fuel load, but native ground cover is adapted to carry a fire in the lightning-season and can burn in a light rain, even when green. 3) Elevated soil moisture can also exist in the lightning-season and there are a many ways to create a patchy burn pattern. 4) Native species of both plants and animals are adapted to cope with fire in the lightning-season; although rare, a few individuals may be injured but there is not date that indicates that mortality is higher and these fires may result in better habitat conditions in the long-run. 5) Unfortunately lightning-season burns may be more uncomfortable for workers; planning helps. 6) Severe burns can occur in any month of the year; outcomes depend on actual fire conditions. ______

Table 3. Some potential benefits of including lightning-season burns in a prescribed fire regime.

1) Bolting is often enhanced in grass-stage longleaf pines. 2) Wildlife food may increase, including arthropods in late summer months. 3) Many species of native ground cover plants will have elevated flowering and seed production. 4) Percent topkill and complete mortality of hardwood stems should increase. 5) Natural ecotones between upland pine forests and adjoining wet areas may be created. 5) There will be an increased number of days for fire each year. 6) There will be a greater range of variability in burning conditions. ______

Table 4. Some questions to address before re-introducing lightning-season fire to a site.

1) What are the long-term management goals for the site? 2) What benefits might be gained by adding lightning-season fires? 3) What is the current fuel load; is there litter buildup under large trees? 4) What is the current vegetation structure; is it close to desired condition? 5) Are some hardwood stems too large to kill with fire? Will they hinder prescribed burning? Are other, short-term techniques needed (cutting, girdling, herbicide) that won’t harm remnant native plants? 6) Are there exotic species on site? How will they respond to lightning-season fire? 7) Will there be differences in wind patterns in the lightning-season compared to past dormant season burns? Will there be different or additional smoke management issues? 8) Are there parties who should be notified and/or educated before shifting season of burn?

58 BURNING ISSUES: REGIONAL PRESCRIBED FIRE COUNCILS PROVIDE SOME SOLUTIONS

David J. Brownlie (U.S. Fish and Wildlife Service, Southeast Region, Fire Ecology Field Office, Tallahassee, FL) Frank T. Cole (For Land’s Sake; Sustainable Forest Management, Wildland Firefighter Foundation, Thomasville, GA)

INTRODUCTION The decline of the longleaf pine ecosystem in the southeastern United States from an estimated 90 million acres when Europeans first arrived, to less than 3 million acres currently, has attracted considerable attention from a diverse range of conservation interests. Declines in many native plant and animal taxa associated with the longleaf pine ecosystem account for much of this interest. The very existence of the Longleaf Alliance, and the attendance at this and prior conferences is testimony enough. There are many government, private, and non-governmental examples of that conservation interest, such as:

z USDA, Forest Service--Longleaf Pine Ecosystem Restoration research program initiated (1993). z U.S. Fish and Wildlife Service--Partners for Fish and Wildlife-Longleaf Pine Ecosystem Restoration Initiative. z Joseph.W. Jones Ecological Center at Ichauway, Newton, GA. z The Nature Conservancy--multiple initiatives.

The pyrogenic, fire dependent nature of the longleaf pine ecosystem, and critical role that prescribed fire plays as a restoration and management tool is also already well documented. A few of the more notable examples include: the 18th Tall Timbers Fire Ecology Conference Proceedings (1993); the Longleaf Alliance Conference Proceedings (1997, 1998) and Longleaf Pine Bibliography (online at http://www.forestry.auburn.edu/la/biblio.htm); and the E.V. Komarek Fire Ecology Database (online at http://www.ttrs.org).

Despite the critical importance of prescribed fire, using this tool has proven consistently problematic for fire practitioners, landowners, and land managers alike. Conflicts between the ecological imperative to restore ecosystem function and biotic diversity and the human imperatives to meet the demands for land and natural resources from an expanding population generate bio-political issues. These bio-political issues limit and threaten prescribed fire use.

THE BURNING ISSUES A brief review of the proceedings from the 20th and 21st Tall Timbers Fire Ecology Conferences (1998 and 2000) yielded the following list of issues and problems confronting prescribed fire users following several decades of relatively effective fire exclusion:

Smoke/Air Quality Backlog-Acres needing treatment Risk of escapes, and erratic fire behavior Wildland-Urban Interface Professional liability/litigious society Gaps in fire effects knowledge base Public understanding/trust Fragmented land ownership and landscape patterns

As part of a survey of prescribed fire practitioners from the eastern United States for a paper prepared for another conference, Engstrom and Brownlie (in press), asked survey respondents to describe the primary barriers to increasing their prescribed fire use. Of 60 responses received from federal, state, and non- governmental (The Nature Conservancy) prescribed fire practitioners in 24 eastern states, 50 (83%) identified barriers. These were grouped into the following categories, listed with the frequency with which respondents cited that barrier:

Proper fuel/weather conditions-too few days (33-34%) Urban interface/fear of escape (18-21%) Too few trained/qualified/experienced people (26-29%) Fire history/effects knowledge (16-18%) Resource objectives conflict/unclear (29%) External barriers (17%) Organizational (internal) barriers (26-28%) Getting desired fire characteristics (14%)

59 Although differing in terms of geographic and demographic focus, and the frequencies with which they were cited, the poster display at this conference (Knott), yields a similar listing of perceived barriers.

REGIONAL FIRE COUNCILS A number of regional fire councils have organized throughout North America, but especially in the southeastern United States, in part to more effectively address many of these bio-political issues. Florida was an early focal point for fire councils and where there are now three councils, along with the southwestern United States. However expanding interest in fire councils continues, as evidenced by the Georgia Prescribed Fire Council, South Mississippi Prescribed Fire Council, Southern Michigan Prescribed Fire Council, and the Interior West Fire Council (which includes several provinces in Canada). Most fire councils are comprised of both government and non-government members, and private landowners that use, or are in other ways affected by wildland fire, especially prescribed fire Miller (1998) provided a summary of the circumstances and early history of the three Florida fire councils, where necessity bred innovation. Both the North and Central Florida Prescribed Fire Councils share the same mission and objectives, whose basic elements are shared by most councils.

Missions: “Encourage the exchange of information, techniques, and experiences among practitioners of prescribed fire in Florida, and to promote public understanding of the importance and benefits of prescribed fire.” Objectives: (from Bylaws): z Provide a framework for communications in relation to prescribed fire objectives, techniques, and issues; z Review prescribed fire problems and develop courses of action; z Disseminate technical information; z Promote the development and utilization of prescribed fire practices commensurate with desirable environmental resource management z Promote public understanding of the benefits of prescribed fire.

The councils pursue their respective missions and objectives in a variety of ways. Nearly all initially organized around, and remain active in legislative and regulatory arenas. For example, all three Florida councils worked to influence state legislation to prohibit prescribed fire use statewide, securing changes restricting the prohibition to just a few counties, and ultimately defeat of the legislation. The Georgia Prescribed Fire Council recently confronted a similar proposal that would have expanded a 13-county (1.2 million acre), May-September annual outdoor burning ban in the Atlanta metropolitan area, to a 45-county (5.5 million acre) area. Only 6 counties were added, and a provision reserving the right to use understory prescribed burns during that period in the other 26 counties was secured in the final version. Perhaps far more important, are pro-active legislative initiatives undertaken by different councils. For example, an initiative to secure formal resolutions and/or ordinances supporting the use of prescribed fire from every county in Florida is nearing completion after several years, and was a vital unity-building endeavor for all three Florida councils. Councils in Florida and Georgia have also secured increased liability protection and wider prescription “windows” for fire practitioners (and indirectly landowners) who voluntarily secure state certification, which in turn recognizes practitioners’ professional status. These statutes are now used as models by other states, even where there are currently no active councils. Florida council efforts led to a 1997 proclamation of the now annual, Florida Prescribed Fire Awareness Week. The Georgia legislature recently funded a new Prescribed Fire Coordinator position within Forestry Commission, specifically to promote prescribed burning, institutionalizing prescribed fire as a formal role for the Commission for the first time.

Perhaps more important than legislative-regulatory activities, are information transfer, and educational outreach activities. Most councils hold steering committee and full council meetings 1-2 times annually. Meetings typically include talks and demonstrations aimed at helping practitioners stay current on applicable laws, regulations, emerging issues, training and continuing education opportunities. The “Lessons Learned/What Went Wrong” segment (modeled after the military’s After Action De-briefing system) has become especially popular with the membership of several councils. For some of the newer councils, these meetings are also crucial for internal council organizational and strategic planning, as undertaken by the Georgia, South Mississippi, and Southern Michigan Councils. Several councils have their own internet

60 websites on which fire training databases, public informational or educational materials (brochures, flyers, etc.), a calendar of events, as well as links to other sites/sources of fire information are posted. The Central Florida Prescribed Fire Council has hosted six “Media Academy” days, specifically addressing prescribed fire that have proven very popular with television, radio, and print media. The North Florida Prescribed Fire Council has their first Media Academy scheduled for late January 2001. Council members present informational talks to civic organizations, regional/county/city planning commissions and other groups promoting the value of prescribed fire whenever invited as an informal “speaker’s bureau”.

Fire council members also cooperate with neighboring councils sharing ideas, products, and other information in a non-competitive/non-proprietary manner. This has been especially beneficial during the early establishment period for new councils. When the Southwest Interagency Fire Council was first forming, help from some of the original “pioneers” in prescribed burning in the southeast was solicited (Murphy and Cole 1998). Florida council members helped core groups of interested people establish the Georgia council. Subsequently Florida and Georgia council members continue assisting the Southern Mississippi Council. More recently, a Florida council member spoke to a southern Michigan group interested in starting a new council, a process watched closely by others from northern Michigan.

CONCLUSIONS Small Bites & Big Ideas--Regional fire councils have effectively addressed many of the bio-political issues emerging from and surrounding prescribed fire use, a vital tool in restoring and maintaining healthy, functioning longleaf pine ecosystems. A need for additional, councils remains in large portions of the longleaf ecosystem’s range, most notably from Alabama west to east Texas, and in the Carolinas. The collective experience of existing councils offer a few basic principles for successful establishment and effectiveness of new councils, perhaps best summed up as “Small Bites and Big Ideas”.

Taking small, discrete initial actions helps bring a sense of focus and often a sense of urgency to the normally small, initial core group of what must be devoted individuals. Often, there has been a single, common issue with a clear geographic scope (an entire state may be too large) that initially brings together and unifies this core group. It is also imperative for continued long-term success, for the council to engage larger, private landowners-managers as active council participants as soon as possible. This may be the most important and most difficult task facing a new council. Membership and leadership change over time is a given. However, it is imperative that a council maintains a responsive information/contact person at all times, especially during leadership changes. Sustaining momentum and membership interest over the long term requires council leadership to maintain a long-term, broad vision that integrates a cooperative rather than competitive approach. To maintain that vision, council leadership needs to periodically reflect on the perseverance of prescribed fire’s “heroes” such as H.H. Chapman, Herbert Stoddard, Henry Beadel, Ed and Roy Komarek, and Leon Neel, once branded as villains.

ACKNOWLEDGEMENTS This story is really about more contemporary heroes, which would not have been told without generous contributions from those now carrying the flame. The authors extend their appreciation to Jeff Caster, Lane Green, Jim Karels, Steve Miller, and Jim Stevenson from the North Florida Prescribed Fire Council, Jim Durrwatcher from the South Florida Fire Council, Geoff Babb, Mark Hebb, and Butch Neal from the Central Florida Prescribed Fire Council, George Ramseur and Tony Wilder with the Southern Mississippi Prescribed Fire Council Steering Committee, and Alan Dozier and Bob Farris with the Georgia Prescribed Fire Council.

LITERATURE CITED Engstrom, Todd R., and David J. Brownlie. [in preparation]. Burning for birds: Concepts and applications. Symposium on The Role of Fire for Nongame Wildlife Management and Community Restoration: Traditional Uses and New Directions at The Wildlife Society, 7th annual Conference, September 12-16, 2000, Nashville, TN. Hermann, Sharon M. (ed.) 1993. The longleaf pine ecosystem: Ecology, Restoration, and Management. Tall Timbers Fire Ecology Conference Proceedings. No. 18. Tall Timbers Research Station, Tallahassee, FL.

61 Kush, John S. [ed.] 1997. Longleaf Pine: A regional perspective of challenges and opportunities. Proceedings First Longleaf Alliance Conference., Mobile, Alabama, September 17-19, 1996. Longleaf Alliance Report No. 1, Auburn, AL. 178 p. Kush, John S. [ed.] 1998. Longleaf Pine Ecosystem Restoration Symposium: Ecological Restoration and regional conservation strategies. Presented at the 9th annual international conference of the Society for Ecological Restoration, November 12-15, 1997, Fort Lauderdale, FL. Longleaf Alliance Report No. 3, Auburn, AL. 87 p. Miller, Steven R. 1998. Florida’s regional fire councils: tools for fire management. Pages 41-43 in Teresa L. Pruden and Leonard A. Brennan (eds.). Fire in ecosystem management: shifting the paradigm from suppression to prescription. Tall Timbers Fire Ecology Conference Proceedings, No. 20. Tall Timbers Research Station, Tallahassee, FL. Moser, W. Keith and Cynthia F. Moser (ed.). 2000. Fire and forest ecology: innovative silviculture and vegetation management. Tall Timbers Fire Ecology Conference Proceedings, No. 21. Tall Timbers Research Station, Tallahassee, FL. Murphy, James L., and Frank T. Cole. 1998. Villains to heroes: overcoming the prescribed burner versus forest firefighter paradox. Pages 17-22 in Teresa L. Pruden and Leonard A. Brennan (eds.). Fire in ecosystem management: shifting the paradigm from suppression to prescription. Tall Timbers Fire Ecology Conference Proceedings, No. 20. Tall Timbers Research Station, Tallahassee, FL. Pruden, Teresa L. and Leonard A. Brennan (eds.). 1998. Fire in ecosystem management: shifting the paradigm from suppression to prescription. Tall Timbers Fire Ecology Conference Proceedings, No. 20. Tall Timbers Research Station, Tallahassee, FL.

ADDITIONAL INFORMATION/CONTACTS:

North Florida Prescribed Fire Council James Furman, Chairman: 850 882-4164, ext. 312 Eglin AFB, AAC/EMSNP 107 Hwy. 85 N Niceville, FL 32578-5321 [email protected] Central Florida Prescribed Fire Council Butch Neal, Chairman: 407 647-7275 PBSJ, Inc. 1560 Orange Ave Winter Park, FL 32789 [email protected]

Georgia Prescribed Fire Council Randy Tate, TNC, [email protected], 404 873-6946 Bob Farris, GFC, [email protected], 912 751-3492

Southern Mississippi Prescribed Fire Council George Ramseur, TNC, [email protected] 228 872-8452

Internet Site: http://flame.fl-dof.com/Env/councils.html

62 PRACTICAL APPLICATIONS OF PRESCRIBED BURNING ON THE CONECUH NATIONAL FOREST

John R. Lint (Calcasieu Ranger District, 9912 Highway 28 West Boyce, LA 71409; previously at Conecuh National Forest, Rt. 5 Box 157, Andalusia, AL 36420) Gary L. Taylor (Conecuh National Forest, Rt. 5 Box 157, Andalusia, AL 36420)

In the longleaf ecosystem, the understory is vital to its biodiversity. Continuous disturbance from fire is the key to a diverse understory. Lightning fires have historically been a primary selective force in maintaining plant and animal biodiversity in these ecosystems (Robson and Myers 1992). In addition, fires caused by Native Americans also influence biodiversity. Longleaf pine is adapted to fire. From its germination on soil exposed by fire, to its thick bark, the longleaf pine has developed numerous attributes that allow it to thrive in areas subject to frequent fires (Stout and Marion 1993). Like the need for physical exercise for maintaining a healthy body, fire improves the fitness of the longleaf pine ecosystem, making it more resilient to insects/disease, and a more productive forest. In other words, “oxidation” (i.e., burning) is the ultimate aerobic exercise for the longleaf pine ecosystem. Much of our forests however, have generally spent most of the last century in poor health, getting soft and out of shape. In many ways brush growth in the absence of fire (e.g., fuel loading) is similar to cholesterol clogging arteries. The brush blocks energy flow in the longleaf pine ecosystem. The worse case scenario from this result is a “heart attack” as seen when a wildfire occurs. Prescribed burning is not a magic pill, but it provides renewed rehabilitation and an ecosystem fitness program.

Not too long ago, much of what is now more commonly accepted in the use of prescribed fires was still relatively new and, in many cases, even considered “taboo.” Ideas such as burning young longleaf stands, burning large areas in the growing season, and subsequently scorching pine stands were thought to be extremely detrimental to longleaf and wildlife. Beginning around 1990, we set out to double the prescribed burning program on the Conecuh National Forest and “re-introduce” growing season burning. This was similar to what many managers were doing across the longleaf pine range at about the same time.

The Conecuh National Forest is composed of approximately 84,000 acres in the coastal plains of Alabama. Longleaf pine dominates over 63,000 acres of pine and pine-hardwood stands. The remainder is hardwood and hardwood pine. Historic accounts in the southeast describe an open, park-like, fire-maintained ecosystem dominated by longleaf pines with an understory of fine grasses (Bartram 1791, Harper 1914). Before establishment as a National Forest in 1936 and similar to the rest of the longleaf range, the Conecuh was cut over and abandoned. Efforts to replant the forest began immediately with the Civilian Conservation Corps in 1936. While the former cut over landscape is no longer evident on the Conecuh, the goal of ecosystem restoration continues.

Goal – Ecosystem Restoration Since the late 1950s, the Conecuh had a history of burning on a 4-8 year rotation with relatively "cool" winter burns and was in reasonably “good shape.” However, we were not satisfied with maintaining the status quo. To fulfill our goal of ecosystem management, we needed to integrate as many resources as possible. We needed to manage for rare species such as the red-cockaded woodpecker, the gopher tortoise, and rare plants as well as manage for popular game species such as white-tailed deer, wild turkey, and bobwhite quail.

Objectives - Increase prescribed burning frequencies from 4-8 years to 2-4 years. This involved increasing from only dormant season burning 12-16,000 acres/year to a 50:50 dormant and growing season mix of 22-24,000 acres/year. - Recognize and deal with our cumulative burning deficit. The fuel loading (amount of burnable material present) continuously builds. Without burning, the fuel type also will change from finer, grassy fuels to more dense brush. Over time, an area that misses burning rotations will require increased burning frequency and/or increased burning intensity to deal with this burning deficit. - Learn more about growing season burning as we increased from 350 acres in 1991 to 12,000-14,000 acres by 1996.

63 - Integrate wildlife, endangered species, timber, and recreation resource areas under the umbrella of ecosystem management.

WHAT WE LEARNED On average, we could not consistently rely on dormant season burning to produce the most effective results. This resulted, over many years, in relatively low-intensity burns that could not reduce woody brush and actually tended to promote brushy conditions. The idea of “top kill” actually increased the existing brush by fertilizing it and the subsequent numerous sprouts. With growing season burns, we could convert fuel types from heavier brush to more manageable lighter, flashy fuels such as grass. Smoke concerns lessened because of this. Without the heavier fuels burning into the late evening, nighttime down-drain smoke decreased. By reducing heavy fuels surrounding the base of trees, mortality decreased.

We also learned that by burning in the dormant season first, then performing a burn in the growing season 1½ years later would reduce the severity that could occur when introducing growing season burns. Many concerns led back to the fuel type and loading. We could burn young longleaf pine stands within 2 years of planting if we kept the fuel loading low. If the site preparation burn was effective, then we could burn as soon as fuel would carry fire. We observed that hardwood mortality such as dogwoods largely depends on fuel loading. If we could reduce the fuel loading with a winter burn and quickly introduce a growing season burn, then dogwood mortality was not as severe. Less frequent and less effective burns produced a cumulative deficit in the fuels, which could be overcome in the next few years by increased frequency and more intense burns. Growing season burning was more effective in changing fuel types to make up for the deficit.

The land, the habitat quickly responds; however, positive results would not last more than 2-3 years. In other words, we could not achieve our goals with just one prescribed burn. We had to introduce a burn to deal with the fuel loading (usually dormant) then perform several growing season burns to convert the fuel type. The quick response and recovery of the longleaf and the understory reinforced that the longleaf is well adapted for frequent burning. Surveys suggested positive results from the prescribed burning on birds (Hill 1998) and pitcher plant sites (Sorrie 1997).

The essential factor in our prescribed burning program was our workforce. There were many unknowns, and often we were proposing things such as growing season burns, larger burn blocks, more frequent return intervals that were different from standard, accepted practices. It was critical for people to get together, use the best science available, and begin trying different techniques. We would not have been successful without the involvement and support of our workforce. Additionally, we involved partners from Auburn University, the Longleaf Alliance, the Nature Conservancy, other state and federal agencies, and the public. As we set out to learn, it was critical to remember that when something went wrong or did not work as expected, we had to look beyond circumstantial proximate causes. For example, if there was tree scorch or mortality, was it the season of the burn, or was it the accumulation of fuel load and the heavy fuel type?

CONCLUSIONS The hardest part seemed to be getting started, but the more burning we did, the easier it got. We began with developing a clear vision of where we wanted to take the Conecuh. We developed goals and objectives. A critical ingredient was to get people together (both Forest Service and our partners) and share information. There were no cookbook answers. We found that the best objectives revolved around making our efforts a learning process with one basic tenet that fire was foundation to the longleaf pine ecosystem.

There is no substitute for the role of fire in the longleaf ecosystem. Fire is a necessary part of the life cycle of many plants and animals as well as the primary disturbance regime that drives the ecosystem. Timber harvest actions (such as thinning, restoration cutting, and uneven-aged management) are important methods of restoring land that was once cut over, fire-excluded, and, then later, primarily winter-burned. These silvicultural treatments could not be fully successful without employing prescribed burning. Likewise, the prescribed burning cannot be fully successful in dealing with the brushy, dense stands with little herbaceous ground cover without performing the needed silvicultural treatments.

64 Prescribed burning is the most effective and cheapest management tool we have for restoration, game, and non-game, endangered species, and hazard fuel reduction. How objectives are accomplished depends on scale, work force, and complexity possible. Objectives are dependent on what you can do matched with what you want to do.

Lastly, you cannot fool Mother Nature. Eventually the woods will burn. We witnessed the differences between wildfire severity during extreme drought years from areas that had received prescribe burns to lower the fuel loading to areas that had heavy fuel accumulations. Our goal was to capture the benefits of fire without the devastating effects that uncontrolled wildfires can bring. Longleaf pine is the premier fire adapted tree we have; it survives and supports fire the best. When increasing a burning program or shifting from dormant season burning to growing season burning, the fire will be more intense and the changes more severe. Those aches and pains should pass as the woods get in better shape. To learn more about getting fire on your land, and to learn how to get more from a dynamic, managed forest, then contact the Longleaf Alliance and local agencies.

ACKNOWLEDGEMENTS The progress and accomplishments in this report were the result of the teamwork and dedication of the personnel of the Conecuh National Forest. They took on the challenge of burning in the hottest time of year. This work was accomplished with advice from our partners: Kenneth Johnson and Thagard Colvin, Alabama State Game and Fish; Vernon Compton, the Gulf Coastal Plain Ecosystem Partnership; Dr. Craig Guyer and Dr. Geoff Hill, Auburn University, Rhett Johnson and Mark Hainds, The Longleaf Alliance, and Mark Bailey, Alabama Natural Heritage Program. We also thank Dr. Sue Grace, USGS for help with this manuscript.

CITATIONS Bartram, W. 1791. Travels through North and South Carolina, Georgia, East and West Florida, the Cherokee Country, the extensive territories of the Muscogulges or Creek Confederacy, and the country of the Choctaws: containing an account of the soil and natural productions of those regions, together with observations on the manners of the Indians. Philadelphia: James and Johnson. 414 pp. Harper, R.M. 1914. Geography and vegetation of northern Florida. In 6th Annual Report of the Florida State Geological Survey, Tallahassee, pp. 71-307. Hill, G. 1998. Use of forested habitat by breeding birds in the gulf coastal plain. So. J. of Appl. For. 22:133-137. Robson L.E. and R.L. Myers. 1992. Seasonal effects of prescribed burning in Florida: a review. Tall Timbers Res., Inc. misc. Publ. No. 8. 96 pp. Sorrie, B.A. 1997. Alabama and Mississippi seepage bog survey: final report, Alabama Natural Heritage Program and USDI Fish and Wildlife Service. 17 p. Stout, I.J. and W.R. Marion. 1993. Pine flatwoods and xeric pine forests of the southern (lower) coastal plain in W.H. Martin, S.G. Boyce, and A.C. Echternacht (eds.), Biodiversity of the Southeastern United States: lowland terrestrial communities. Wiley and Sons, Inc. New York. pp 373-446.

65 INTERAGENCY COOPERATION AND THE NATIONAL INTERAGENCY PRESCRIBED FIRE TRAINING CENTER

John Fort (NIPFTC, Tallahassee, FL)

Much has been made of the need for interagency cooperation in the field of fire management. Indeed, it is commonly acknowledged, nationally, that fire management agencies can no longer individually carry out the financial, policy, and public outreach initiatives that today’s programs dictate. (It is to be noted that fire “management”, not fire “suppression or control”, is deliberately the term of choice. The days of viewing wildland fire management as no more than “putting them out quickly” are over.) Financially, public demands assure that no single agency has the budget to implement the level of protection (suppression), and enhancement (ecosystem health) that is presently being requested. A common position relative to wildland fire policy will never be totally amenable to all resource management agencies due to the divergent publics each is responsible to. However, in the last five years steps have been taken in the federal sector to unify policy, and federal cost share funding guidelines are facilitating a common policy among the state and private sectors. Relative to outreach, it is counterproductive for each agency to approach the public with less than a unified voice; confusion and mistrust will soon follow.

Fire Management has made strides in the past twenty years to unite efforts. In earlier years it was not unusual for the U.S. Department of Agriculture to speak with one voice, the U.S. Department of Interior to speak with another, and the State and Private sectors to, frequently, not have a voice at all. Times have changed, as evidenced by several diverse initiatives.

Twenty years ago agencies began to speak with a common vernacular, the Incident Command System, or ICS. While initially viewed as the Incredible California System (where it originated), ICS has proven to be an amazingly versatile and adaptable method for diverse agencies to communicate on the projects they implement. Agencies now have a common language. In the last fifteen years coordination efforts have become unified. The dispatch (coordination) centers across the country are staffed by the agencies encompassed by a center’s geographic area; no more individual centers for individual agencies. Indeed, states now play a role in these centers. In the last ten years Fire Councils (or their semantic equivalent) have arisen across the country. Originating in Florida, and spreading to the west coast, these are loose confederations of fire management professionals, concerned citizens, and interested parties that regularly and informally meet to discuss and address topics of common concern and interest. These councils can, and do, affect policy. In the recent past, the federal government (fire management’s deep pocket) has taken fiscal steps to lower interagency barriers. Agencies are now directed not to cross bill each other for fire incidents/projects that are worked on jointly. A fairly subtle policy, but one that has meant a marked improvement in field level relations. Even now the climate continues to improve; the Department of Defense has become signatory to NWCG, the national interagency coordinating group for wildland fire.

Are these efforts adequate to today’s fire management spectrum? They are the basis for a continuing evolution. Future progress can be made in several areas. Most significantly, a common outreach is needed nationwide. If one common denominator could be identified, it would be public information and education. An informed public would serve as the foundation for intelligent fire management on our national wildlands. What is needed is a program on the level of Smokey the Bear’s overly successful message of the last fifty years. A new national initiative must be chartered to speak to the demands of our present degraded ecosystems. If only a fraction of the dollars presently being expended on suppression was dedicated to a professional, deliberate, and unified interagency information campaign, more long term benefit could be realized. For years the public united behind Smokey. They would again unite, if they were informed of and aligned with efforts to restore the health of fire dependent ecosystems.

Internal agency administrators, private, state, and federal, must become amenable to large scale, ecosystem and landscape based fire management plans. These plans in the future will routinely cross administrative boundaries, commonly addressing landscape and political realities. These plans will have the blessing of an informed public and will treat landscape and drainage scale acreage. This will become an interagency reality. When it will happen will be determined by how soon national leadership will provide for a broad

66 based management philosophy, as opposed to a suppression mentality. And the field orientation towards fire management as little more than a source of adrenaline and overtime must change as well. One final observation, which may be viewed by a retentive few as heresy, deals with the age-old concept of a single unified national natural resource agency. While this agency is politically far from reality, a unified national fire management agency may be closer, and warrants real, measured discussion.

What role does the National Interagency Prescribed Fire Training Center (NIPFTC) play in the above discussion? It is a pivotal role. NIPFTC as a concept dates to the early nineties, when it was batted about campfires and beers on the Apalachicola National Forest. In the mid nineties the idea was tried by inviting teams of fire professionals from the western states to work on prescribed fire projects in the Tallahassee area. The response was positive, and in 1997 national offices of the U.S. Forest Service and U.S. Fish and Wildlife Service provided seed money to establish a unique national center of excellence in Tallahassee. This Center was, and is, charged with developing and implementing a field oriented, experience based program. The charge is to provide both national and international fire managers with the focus and foundation to treat fire adapted communities with a rational, broad scale approach. The Center’s main approach is to “learn by doing”.

Classroom and formal coursework relative to fire is available anywhere in the country. What is not readily available is a practical, outdoor avenue to witness and implement the many facets that compose the art of fire management, specifically prescribed fire. In Florida, and the Southeast in general, fire management in the twentieth century was never relegated to a backseat, boogyman role. Broad scale prescribed fire has a long history in the Southern piney woods as a desirable tool to lessen wildfire occurrence and intensity. As well, this same prescribed fire has long been used for ecosystem enhancement. Well established, generations old programs abound in this part of the country. And the managers of these programs are willing to share their successes. And failures. Program management is central to NIPFTC’s vision. Program management shares common denominators regardless of location or ecosystem.

The center functions as a conduit, bringing together established, broad scale programs. These programs are then tied to practitioners eager to learn and implement broad fire management initiatives. The practitioners are allowed to learn by doing, in an environment that encourages deliberate action and observation of effects. These folks operate in six person teams, led by an experienced coordinator, and are self-contained and mobile. The length of stay is twenty- two days. Individual teams are diverse, composed of participants from different agencies, locations, and experience levels. Every effort is made to provide each team with burns in differing fuels and with differing host agencies. Participant confidence is increased and a broad scale interagency approach to fire management is facilitated.

The Center has been widely acknowledged as successful. It continues to grow and innovate. Recent national increases in hiring and in levels of federal funding insure a continuing need for individual and program development. Information on the center and its accomplishments is available at:

http://fire.r9.fws.gov/pftc

or by calling (877) 223-2198

67 SOUTHEASTERN ECOLOGICAL FRAMEWORK

Stephanie Fulton (US EPA/OPM-Planning, 61 Forsyth Street, Atlanta, GA 30303)

What is an Ecological Framework? The natural environment and the processes that support it are our life support system. Every thing that the environment provides to us for free, usually comes at a very high price if we have to replace it or maintain it. In that regard, preservation of existing natural systems and their inherent processes is essential for our survival. Landscape ecologists have known for a long time that piece-meal protection of the environment often leads to degradation of the parts being protected. The resulting fragmentation prevents the operation of many large-scale processes from adequately functioning. Preservation of natural areas that are contiguous with other natural areas is an important principle.

The Southeastern Ecological Framework is a prototype for the protection of water quality, species habitat, important ecological areas, quality of life and other important natural features by protecting large, intact landscapes and connectivity between such areas. The Southeastern Ecological Framework follows natural land and water features such as rivers, ridges, estuaries, wetland basins and upland forests at a regional (8 state) scale. The framework is comprised of important regional ecological hubs and corridors that connect them. The hubs of the framework are typically land areas with important riparian areas, no or few roads, high habitat diversity, little habitat fragmentation, rare habitats or species, and greater than 5,000 acres in size. Often they are associated with existing managed lands such as wildlife refuges, parks, national forests or private protected lands. The corridors of the framework connect the hubs and typically follow natural land forms and water features, allowing ecosystem processes to operate at a larger scale.

What is this project about? With funding from the Intermodal Surface Transportation Efficiency Act (ISTEA), the University of Florida developed a model to identify potential greenways and trails in Florida. That project developed the modeling protocol and expertise for designing landscape linkages and prioritizing ecological hubs. The Florida Greenways and Trails model underwent significant public participation, comment and peer review. The University of Florida Departments of Landscape Architecture, Urban and Regional Planning and Wildlife Ecology and Conservation were awarded a cooperative agreement grant to develop an ecological connectivity model for the eight states in the Environmental Protection Agency’s Southeast Region (4). The purpose of the project is to identify regionally significant lands that would aid in protection of water

EPA’s Mission The mission of the Environmental Protection Agency is to protect human health and to safeguard the natural environment – air, water EPA Region 4 Planning and Analysis Branch and land – upon which life depends.Strategic Environmental Assessment Corps Contact: Source(s):

resources, wetlands, and other natural areas. The conservation of native landscapes and ecosystems also connect people to the land with other archeological, historical and cultural resources. The finished product is

68 a place-based coverage of ecological hubs and corridors in the Southeast. The resulting framework represents some of the best remaining large intact ecological areas in the southeastern states of Georgia, Florida, Mississippi, Alabama, Tennessee, Kentucky, South and North Carolina.

Why is this important? A green infrastructure in the southeast can have significant ecological, economic and social benefits for the region. From an environmental point of view, the Southeastern Ecological Framework can be an important component of regional, state and local conservation efforts. From EPA’s perspective, the models and data can play an innovative and multi-purpose role in protecting water and air quality. The approach can serve other agency missions by contributing to wildlife habitat conservation, prioritizing wetland mitigation locations, sequestering carbon and removing particulate matter, identifying growth management strategies, and more.

This product can offer a wealth of opportunity as a template for other federal and state agencies and non- profits to coordinate programmatic activities that support environmental protection while maintaining ecosystem connectivity. One example of coordination is in direct support of Georgia’s Community Greenspace Program. The Southeastern Ecological Framework is helping local governments prioritize land that connects communities with their natural surroundings. Through voluntary efforts, the fastest growing counties in Georgia are identifying 20% of available greenspace and designing plans to protect this vital resource. EPA Region 4 and the Trust for Public Land are working with county governments to identify local connectivity within the context of a statewide greenway network.

What can it be used for? The Southeastern Ecological Framework is a template of important ecological areas that can be used for many purposes: 1) The development of mitigation banks and sites that provide connectivity to larger intact wetland systems, 2) buffering of protected wildlife or natural areas such as wildlife refuges, national parks, state and local parks and private wilderness areas, 3) planning of future road right of ways to minimize impacts on existing natural areas, 4) watershed protection and guidance for siting of future industrial activities, 5) prioritization for areas in conservation through wetland reserve or conservation reserve programs, 6) siting for reforestation of riparian areas, 7) integration of local greenspace protection efforts into the larger regional picture, 8) conservation reserve design and planning for conserving biological diversity. This list is only a start of the types of activities that can be planned around the SE ecological framework to preserve and protect our dwindling natural resource base.

Who to contact: Rick Durbrow, Program Analyst, US EPA, 404-562-8286 [email protected]

John Richardson, Project Officer, US EPA, 404-562-8290 [email protected]

Tom Hoctor, UFL Dept. of Wildlife Ecology & Conservation, 352-392-5037, [email protected]

Margaret Carr, Principal Investigator, University of Florida Department of Landscape Architecture, 352-392- 6098 ex 327 [email protected]

University of Florida Geoplan website: www.geoplan.ufl.edu/epa/index.html

69 OBSERVATIONS AND THOUGHTS FROM THE PANEL DISCUSSIONS

ECOLOGICAL AND ECONOMIC CONSEQUENCES OF THE 1998 FLORIDA WILDFIRES

Sue Grace (US Geological Survery, National Wetlands Research Center, Lafayette, LA) Dale Wade (US Forest Service, Athens, GA)

EXECUTIVE SUMMARY Over 2,000 wildfires burned 500,000 acres of Florida real estate, most of it between May and mid July 1998. Although virtually every county was impacted, the fires were concentrated in the northeast quadrant defined by boundaries extending north from Orlando to the Georgia line and east from Orlando to the Atlantic Ocean. The fires occurred during record-breaking drought, consumed vast amounts of accumulated fuel in normally wet depressions that rarely burn in prescribed fires or during more typical wildfire conditions, and crowned through pine plantations and subdivisions forcing the evacuation of an entire county. Such extreme fire behavior is unusual but not unprecedented in Florida. Property damage, economic ramifications such as airport closures and tourism losses, natural resource damage, and suppression costs were estimated at between $620 and $890 million, ranking this as one of Florida’s worst disasters. Air quality impacts such as respiratory problems requiring medical treatment were not addressed. The Joint Fire Science Board saw these fires as an opportunity to scientifically test some of the hypotheses raised in the wake of this catastrophe that resulted from the combination of two extreme events – record-breaking drought and an unusually high number of dry lightning storms. The research team assembled was comprised of people from: US Forest Service, Southern Research Station; US Geologic Survey, Biological Research Division; Florida Division of Forestry, Fire Control and Forest Management Bureaus; Florida Natural Areas Inventory; St. Johns River Water Management District; Auburn University; Dynamac Corporation; GP The Timber Company and; The Nature Conservancy. Study sites included a national forest, national wildlife refuge, several state forests, wildlife management areas, a state reserve, a water management district, and industrial woodlands. Individual study topics, objectives, results, and some management implications are presented below. Complete reports on each of these studies including a map of the fires, tables and figures can be found through the Florida Forest Protection Bureau web site at >http://flame.fl-dof.com/joint_fire_sciences<.

1) Topic: Effects of Silvicultural Practices on Extreme Fire Behavior Objective: • Determine potential fire behavior in pine flatwoods following partial timber harvest, prescription fire or understory herbicide application. Results: • For immediate reduction in potential wildfire behavior, prescription fire is best, but because of quick understory recovery, fire must be applied at least every 5 years. • Partial harvest also provides immediate short-lived reduction in potential fire behavior, but necessary return interval is economically impractical. • Eradication of understory with herbicides has no immediate fire behavior reduction benefits because dead stems remain standing. Beginning the 2nd year after treatment, however, potential fireline intensity decreased substantially and remained low for at least 6 years. • Only prescribed fire and partial harvest reduced forest floor accumulations. Management Implications: • A short-interval prescribed fire program will substantially reduce overstory pine mortality during subsequent wildfires. • In plantations a single herbicide application at crown closure will minimize potential fire behavior until harvest after a 2-year lag. • Drought-season fires in herbicide-treated stands will have lower potential fireline intensity, but they will still be high severity and thus kill root systems, resulting in substantial overstory mortality. • Combining fire and herbicide treatments untested, but should provide immediate and long- term reduction in both fire intensity and severity.

70 2) Topic: Effects of Fuel Treatment on Overstory Mortality of Southern Pines Objectives: • Quantify the effects of an array of prescribed fire frequencies on southern pine mortality after wildfire. • Determine the effects of stand origin and site moisture level on overstory mortality after wildfire. Results: • Prescribed burn history significantly affected mortality. • Mortality lowest in stands prescribed burned 1.5 years before wildfire (<10%). • Mortality highest in stands where prescription fire not used (89%). • Mortality significantly higher on normally wet areas than elsewhere (65% vs. 30%). • No difference in mortality on sites burned by headfires and those burned by backfires. • Little delayed mortality in planted stands, but some during 2 year in natural stands. • Crown loss >75% was a good predictor of 2nd year mortality in natural stands. • Stand origin (planted or natural) had no affect on overstory mortality when data adjusted for fact that trees usually not planted in normally-wet depressions.

Management Implications: • A short-interval prescribed fire program will substantially reduce mortality during subsequent severe-drought wildfires. • During severe drought fires, there is no difference in mortality between sites burned with headfires and those burned with backfires. • Mortality is confined to 1st year in plantations but some takes place 2nd year in natural stands. • 2nd year mortality can be reliably estimated 4 months after severe-drought wildfires.

3) Topic: Performance of the BEHAVE Fire Prediction Model Objectives: • Quantify BEHAVE model performance by comparing predicted to observed fire behavior. • Document how Fire Behavior Analysts (FBA’s) operationally used the BEHAVE model. Results: • First objective could not be met because copies of BEHAVE model runs were not found in archived FBA reports, nor was sufficient documentation provided to reproduce their BEHAVE runs. • No single BEHAVE model worked so FBA’s improvised to make predictions agree with observations by using different models for different outputs, adjusting input parameters until predictions matched observed behavior, or my abandoning the use of BEHAVE altogether and relying on their own expertise. • Problems in data collection procedures are described and potential solutions suggested. Management Implications: • FBA’s recognize the shortcomings of BEHAVE models and improvise to make reliable fire behavior predictions.) • BEHAVE archive problems can be fixed.

4) Topic: Predictors of Extreme Fire Behavior Objective: • Correlate daily fire behavior, as measured by rapid fire growth, to the Atmospheric Dispersion Index (Lavdas Index) and the Lower Atmospheric Stability Index (Haines Index). Results: • Both indexes preformed well in predicting large fires in 1998 but not in 1999. • Lack of wind information in Haines Index appeared to be a serious limiting factor. • Lavdas Index had larger number of false alarms.

71 • A new index equivalent to the ratio of the Lavdas stability component to the transport windspeed component was best both years and had the fewest false alarms. Management Implications: • Neither the Lavdas Index nor Haines Index proved to be reliable predictors of extreme fire behavior days in Florida.

5) Topic: Short-term Response of Plant Species of Special Concern and Exotics. Objectives: • Assess the status and response of known populations of plant species of special concern. • Identify and map new populations discovered during the course of the study. • Determine the extent of exotic species invasions on burn areas and in adjacent control lines. Results: • Known populations of species of special concern all appeared to benefit. • New populations of several species were found and mapped. • Benefits were exemplified by the federally endangered Rugel’s pawpaw which increased from 200 to 2000 individuals with increased flowering (80%). • No exotics found on the burns but stable reproducing populations noted nearby. Management Implications: • Growing-season fires prior to August in this ecoprovince are likely to benefit uncommon plant species. • Fresh burns will not necessarily be rapidly invaded by nearby exotic species.

6) Topic: Fragmentation at the Landscape Level Objectives: • Develop GIS-based maps showing fuel conditions useful in assessing fragmentation. • Evaluate GIS-based fire and habitat maps for predicting and interpreting wildfire effects • Determine effects of the fires on habitat suitability and population dynamics on a relatively isolated population of the federally listed Florida Scrub-Jay. • Compare habitat suitability before extensive fire suppression (1943) with changes after 50 years of fire suppression using GIS. • Use GIS to look for differences between areas occupied by Florida Scrub-Jays and unoccupied areas regarding fire history and habitat quality. Results: • Florida Scrub-Jays cannot persist in habitat subjected to infrequent fire. • Atlantic coast populations need more frequent fire than previously believed because vegetation recovers faster near the coast. • GIS maps of landcover and fire boundaries are probably too coarse to predict and interpret the effects of wildfires. • Habitat mapping applications are generally too coarse to provide the information needed for management and predicting population responses. • Forest barriers between occupied and restored habitat should be expeditiously eliminated where the forests are artifacts of human activities. Management Implications: • Good Scrub Jay habitat requires frequent fire. • Coastal scrub needs to be burned more frequently than interior scrub to provide suitable Scrub Jay habitat. • Eliminate barriers between occupied and restored Scrub Jay habitat. • Current GIS map capability too coarse to make specific postfire management decisions and predict population responses.

7) Topic: Insect Responses Objectives: • Determine tree mortality along a fire-intensity gradient. • Monitor abundance of bark beetles and woodborers over time and correlate to tree mortality.

72 • Determine the prevalence of Leptographium spp. fungi in live tree roots. Results: • October 1998 survey showed tree mortality was related to fire intensity and ranged from 2% of trees in unburned control, 9% in low-intensity stands, to 64% in high-intensity stands. • Predicted population explosion of bark beetles and subsequent increase in mortality during the summer of 1999 did not materialize. • Of the trees alive in 10/1998, 31% in the high-intensity stands died, 2% in low intensity stands died, and 2% of the unburned control trees died during the next 12 months. • Very little additional mortality occurred between October 1999 and June 2000 (<3% in high-intensity stands). • No sampled trees in control plots had roots infected with Leptographium spp. • No healthy roots found in high-intensity stands. • >75% of live trees in high-intensity stands infected with Leptographium spp. • 15-20% of sampled roots in moderate- and high- intensity stands had reproduction weevil larval galleries. • 0-4% of sampled roots in controls and low-intensity stands had reproduction weevil larval galleries. Management Implications: • Most tree mortality occurred within the 1st year postfire. • After large intense fires, populations may not explode and attack adjacent unburned trees. • Standing snags may not provide good conditions for brood survival during severe drought periods in FL (hot and dry). • Do not immediately replant or reproduction weevils may be a problem. • Leptographium prevalence is associated with fire intensity. • Root weevil numbers and damage are high in high fire-intensity stands.

8) Topic: Home Protection Strategies Objective: • Evaluate the utility of some commonly recommended home protection strategies. Results: • This study reinforced 1985 study conclusions that use of metal soffits and amount of vegetation clearance around a home were the best homeowner strategies to provide protection from wildfire. • Block construction (but not type of exterior), tile roof, lack of roof and yard debris, and defensible homeowner or fire department actions also all significantly increased the likelihood of home survival. • Wooden privacy fences attached to, and firewood stacked next to, a house increased the likelihood of it sustaining fire damage. • Only 16% of 75 homeowners interviewed were aware of wildfire protection strategies and only 8% had actually implemented one or more protection measures. • The use of prescription fire at the Urban Wildland Interface (WUI) was very effective in protecting homes from subsequent wildfire. Of the 32 treated within 9 months prior to the ’98 wildfires, only one was damaged. Management Implications: • Many of the interviewees were from a subdivision that had also lost several hundred homes to wildfire in 1985. The collective subdivision history of past conflagrations was thus either nonexistent, was not effectively passed on to new arrivals, or was dismissed as a one-time event. • Agency efforts to communicate home protection strategies have not been fully successful and thus need to be increased and have the content and/or delivery system changed. • Subdivision association bylaws, city/county ordinances, or state statutes/rules that spell out minimum structure protection thresholds should be considered.

73 • This analysis reinforces an Australian study that showed allowing homeowners to remain and protect their homes during conflagrations increased the probability of survival of both. • The study shows fuel reduction measures such as prescribed burning at the WUI significantly improve the chance of structure survival during wildfire and should be greatly expanded.

9) Topic: Economic Impacts Objective: • Evaluate the efficacy of fuel reduction treatment policies and programs for reducing the economic impacts of catastrophic forest fires. Results: • Estimates of total damage from the ’98 Florida fires ranged from $622 to $888 million. • In the short run timber prices dropped with the sudden influx of salvaged timber but as soon as that supply declined, prices rose and have remained at a level above what they would have been had the fires not occurred. • Forests at greatest risk under this severe drought were those coniferous stands in or near wetlands, especially bald cypress. • Fragmentation of the forest appeared to increase wildfire risk. • Wildfires within the past decade and reduced understory stature both reduced the likelihood of wildfire in 1998. Prescribed fire on the other hand showed only a weak statistically insignificant benefit. The effects of prescription fire on subsequent wildfire intensity were not addressed. • In prior extreme-fire years, the number of small fires increased with a relative reduction in large fires, but in 1998, there were more large fires relative to small fires. • Urbanization was positively correlated with area burned during 1998 in contrast to previous years. • This ecoprovince accumulated a large wildfire deficit in the nine years prior to 1998. • The 1998 wildfires more than consumed this deficit leaving the region with a wildfire “surplus”. This did not happened in other Florida ecoprovinces during the same time period. Management Implications: • As Florida’s population continues to increase, so will the economic cost of wildfires. • As forests become more fragmented due to urbanization, anthropogenic wildfire risk will continue to increase. • As the wildland urban interface expands, fire control forces are concentrating on structure protection at the expense of fire suppression, which results in larger fires.

74 FIRE IS THE ECOLOGICAL IMPERATIVE IN FIRE-ADAPTED ECOSYSTEMS SUCH AS LONGLEAF PINE

Dale Wade (Opening comments made by at the Fire Ecology Panel)

Fire exclusion invariably results in catastrophic fires such as those if Florida in 1998 and Montana in 2000 directly threaten the health and safety of both the public and firefighters, and result in smoke and haze episodes which in turn aggravate respiratory problems and reduce visibility.

With extended periods of fire exclusion in short-interval fire ecosystems, fuel accumulations reach unprecedented levels resulting in fire intensity that is often outside the range of natural variability and residual smoke problems not encountered under historical fire frequencies.

Thus, after an extended period of fire exclusion, a decision to restore fire by simply allowing wildfires to burn under a generalized prescription is not a viable alternative because such fires can be beyond the natural resiliency of the ecosystem to recover.

In fire-adapted ecosystems, the intentional use of fire is the only choice, in many instances in combination with herbicides or mechanical methods to reduce understory stature.

Decades of fire exclusion in short-interval fire ecosystems cannot be rectified with one or two burns.

These initial restoration burns may have to be conducted during the dormant season to take advantage of low ambient temperatures and uniform forest floor fuel moistures to reduce fireline intensity and forest floor consumption, rather than during the lightning season. Dormant season burns will not restore vegetative communities such as canebrakes to pre-exclusion conditions; higher fireline intensities during the growing season will likely be necessary.

Prescribed fire under site-specific burning conditions such as wind direction will minimize smoke intrusions. A fire prescription must be written to achieve ecological objectives first, and then adjusted as possible to minimize potential smoke problems.

These recommendations are easier stated than executed. As an analogy, one should understand haw a clock works before opening a watch repair shop. This is where research can provide guidance.

75 POSTER SESSION

76 THE ROLE OF RHIZOPOGON MYCORRHIZAE IN DROUGHT TOLERANCE OF LONGLEAF PINE SEEDLINGS

Michael P. Amaranthus (College of Forestry, Oregon State University, Corvallis Oregon, USA 97331) Nicholas Malajczuk (Murdoch University, 14 Metz Way, Wembley Downs, West Australia Australia 6019)

ABSTRACT: Nursery inoculation of mycorrhizal fungi selected for their ability to persist in the outplanted environment and promote drought tolerance could be an important tool for longleaf pine forest establishment and performance. This experiment was conducted to test the ability of Rhizopogon-inoculated and non- inoculated container grown longleaf pine seedlings to tolerate drought conditions over a range of declining soil moistures and increasing soil pressures. As soil moisture levels decreased below 21% the comparisons of plant moisture stress levels between Rhizopogon-inoculated and non-inoculated seedlings were significantly different at the p<0.05 level. Plant moisture stress averaged 5.2 bars for Rhizopogon-inoculated seedlings and 9.3 bars for non-inoculated seedlings at 11% soil moisture. Plant moisture stress averaged 8.2 bars for Rhizopogon-inoculated seedlings and 12.7 bars for non-inoculated seedlings at 7% soil moisture. Plant moisture stress averaged 9.8 bars for Rhizopogon-inoculated seedlings and 18.3 bars for non-inoculated seedlings at 4% soil moisture. Plant moisture stress levels averaged 78% higher for non-inoculated seedlings over the three lowest soil moisture periods. Results from this study indicate that Rhizopogon mycorrhizae can help pine seedlings tolerate soil water deficits.

INTRODUCTION In the southeast United States container-grown longleaf pine seedlings are planted on reforestation sites in increasing numbers. Sandy soils, low water holding capacities, warm temperatures and infrequent spring and early summer precipitation are common across the planting range of longleaf pine . First-year mortality of these seedlings can be high on harsh sites and plant moisture stress is probably the single most important cause. Competing herbaceous vegetation can adversely affect water relations of newly planted conifers (Preest 1977, Peterson and Newton 1985). Transpiration potential during the 3 to 5 months of summer drought can exceed the water storage capacity of the surface soil and herbaceous species can deplete stored soil water early in the growing season, killing or reducing growth of seedlings (Johnsgard 1963).

Conifer seedling establishment depends on rapid root and ectomycorrhizal formation on sites difficult to regenerate (Amaranthus and Perry 1987, 1989a, Mikola 1970). Ectomycorrhizae are commonly assumed to enhance water uptake by their hosts (Trappe and Fogel 1977; Reid 1979; Parke et al 1983). Tolerance to low water potentials varies widely however among mycorrhizal species (Mexal and Reid 1973; Theodorou, 1978; Parke et al. 1983). Theodorou and Bowen (1970) observed that Pinus radiata seedlings inoculated with Rhizopogon lueoulus survived a particularly dry summer better than non-treated control seedlings. Sands and Theodorou (1978) found the resistance to water flow from the soil through mycorrhizal plants was greater than for non-mycorrhizal plants due largely to differences in effective rooting area. Leaf water potentials of mycorrhizal seedlings were lower than for non-mycorrhizal seedlings. It is well established that net photsynthesis declines in response to plant water stress, mainly owing to stomatal closure that accompanies loss of turgor (Boyer, 1976).

Considerable effort and expense is directed toward site preparation and reducing plant competition for limited soil water at longleaf pine planting sites. Inoculum density and viability of native mycorrhizal fungi are often reduced as of a result of site preparation activities (Wright and Tarrant 1958; Amaranthus et al. 1994, 1996; Dumroese et al. 1998). Amaranthus et al. (1996) found significant reductions in mycorrhizal abundance and diversity, including Rhizopogon species with moderate to high levels of organic matter removal and compaction. Clearly, many disturbed forest sites have reduced mycorrhizal forming potential. Certain mycorrhizal fungi such as Rhizopogon form spores within below-ground fruiting truffles. Spores are not distributed airborne, and require animals to dig up the truffles and spread spores across stands and landscapes via fecal pellets. At the time of outplanting longleaf pine seedlings are generally not colonized or colonized by mycorrhizal fungi such as Thelephora terrestrius not well adapted to the harsh outplanting conditions. Colonizing seedlings with Rhizopogon spore inoculum at the nursery assures the presence and benefits of a valuable mycorrhizal type in the outplanted environment.

77 Nursery inoculation of mycorrhizal fungi selected for their ability to persist in the field environment and promote drought tolerance could be an important tool for longleaf pine forest establishment and performance. This experiment was conducted to test the ability of Rhizopogon-inoculated and non-inoculated container grown longleaf pine seedlings to tolerate drought conditions over a range of declining soil moisture and increasing soil pressures.

METHODS In April, 2000, longleaf pine seedlings grown at GHW Weyerhaeuser Nursery in North Carolina were inoculated with 100,000 spores each of the mycorrhizal fungus Rhizopogon rubescens. Spores were applied as a soil drench following maceration of Rhizopogon rubescens sporocarps for 10 minutes in distilled water. Spore concentrations were determined with a haemacytometer.

Fifteen each randomly selected Rhizopogon-inoculated and non-inoculated longleaf pine containers were removed from container plugs and repotted in 1 gallon containers. Sterile sand (75%) and peat moss (25%) was used to fill the one gallon containers upon transplanting. Soils and seedlings were watered thoroughly at the time of transplanting and never watered again. Three seedlings each randomly selected Rhizopogon- inoculated and non-inoculated longleaf pine in each of 5 sample periods. Sampling occurred 48hrs and 4,7,8,and 9 weeks following transplanting into the 1 gallon pots.

Soil moisture percentage was calculated throughout the study including at the time of determination of plant moisture stress. Soil samples were collected to 10 cm depth in each pot following seedling harvest . Samples were packed in airtight containers and taken to the laboratory for weighing, oven-drying for 24 hours at 105 oC and reweighing.

Pre-dawn moisture-stress readings were taken on longleaf pine seedlings 48hrs and 4,7,8, and 9 weeks following transplanting into the1 gallon pots. The seedlings were randomly selected and were severed at the base and placed in a pressure bomb for plant moisture stress analysis. Three seedlings per treatment were selected.. Root systems of all seedlings decapitated for moisture stress readings were and taken to the laboratory for analysis of root and mycorrhizal tips. Roots were gently washed free of soil and extraneous material and subsampled in three cross sections, 1.5 cm broad, of the entire root systems in upper, middle and lower positions, respectively. All active tips were tallied as Rhizopogon, other mycorrhizal or nonmycorrhizal from characteristics observed through a stereomicroscope (2x by 5x magnification). Rhizopogon determination was based on color, surface appearance and thickness of the mantle, short branching morphology, length and characteristics of rhizomorphs. Rhizopogon rubescens mycorrhizae were creamy white and developed a gradient of yellow and reddish coloration with maturity and upon bruising. The Rhizopogon rubescens mycorrhizae had a two-layered mantle and abundant rhizomorphs developing a compact coralloid morphology with maturity.

Statistical analysis was by ANOVA and Tukey's multiple range test. Residuals from the data on percent mycorrhizal colonization and plant moisture stress were plotted to determine if a lognormal transformation was called for to compensate for lognormally distributed values; this indeed was the case for percent mycorrhizal colonization, so the data were accordingly transformed to produce a relatively normal distribution (Steel and Torrie 1980)..

RESULTS AND DISCUSSION As shown in figure 1, there was no statistically significant difference in plant moisture stress between Rhizopogon-inoculated and non-inoculated at 40% soil moisture levels. Similarly, where soils were nearly saturated with water, a soil pressure level of .01, there was no statistically significant difference in plant moisture stress between Rhizopogon-inoculated and non-inoculated (figure 2). There was no statistically significant difference in plant moisture stress between Rhizopogon-inoculated and non-inoculated at 40% and 21% soil moisture levels. However, as soil moisture levels decreased and soil pressure levels increased the comparisons of plant moisture stress levels between Rhizopogon-inoculated and non-inoculated seedlings were significantly different at the p<0.05 level. Plant moisture stress averaged 5.2 bars for Rhizopogon-inoculated seedlings and 9.3 bars for non-inoculated seedlings at 11% soil moisture and 1 bar soil pressure. Plant moisture stress averaged 8.2 bars for Rhizopogon-inoculated seedlings and 12.7 bars for non-inoculated seedlings at 7% soil moisture and 7 bars soil pressure. Plant moisture stress averaged 9.8 bars for Rhizopogon-inoculated

78 seedlings and 18.3 bars for non-inoculated seedlings at 4% soil moisture and 20 bars soil pressure. Plant moisture stress levels average 78% higher for non-inoculated seedlings over the three lowest soil moisture periods.. Percent Rhizopogon colonization on the inoculated longleaf root systems was significantly higher at 4% and 7% soil moisture due to the widespread senescence of non-Rhizopogon roots (figure 3). Three of the non-inoculated control seedlings had minor colonization by Rhizopogon but was less than 5 percent of the total feeder roots. Thelephora terrestrius was common and abundant on both Rhizopogon-inoculated and non- inoculated seedlings.

Rhizopogon mycorrhizal colonization significantly reduced plant moisture stress on longleaf pine seedlings at low levels of soil moisture and high levels of soil pressure. Rhizopogon has certain features which may be influencing its ability to promote drought tolerance. Many Rhizopogon species produce prolific rhizomorphs in soil; these have been noted and described in synthesized ectomycorrhizae (Massicotte et al. 1994; Molina and Trappe 1994; Agerer et al. 1996) Rhizomorphs play an important role in water uptake and movement in ectomycorrhizal systems (Duddridge et al. 1980; Brownlee et al. 1983; Read and Boyd 1986). Parke et al. (1983) and Dosskey et al. (1990) demonstrate enhanced tolerance to drought stress of Douglas-fir seedlings inoculated with R. vinicolor and attribute this enhancement in part to rhizomorph production and function in water transport. Rhizomorphs also link plant roots and perhaps provide conduits for interplant exchange of water and nutrients (Simard et al. 1997).

Rhizopogon also forms a thick sheath surrounding fine roots. Reid (1979) states that it is unlikely that mycorrhizal sheaths themselves function in water absorption but it is conceivable that hydration of spongy fungus mantles, mycelial strands and gels buffered seedlings from drought stress at lower soil moisture contents. It is interesting that in this study that Rhizopogon-inoculation confers benefits to the longleaf pine even when soil volume is artificially limited in the one gallon pots and hyphal growth is restricted. This suggests that absorption and/or transport by mycorrhizal fungi is more effective than by roots. In many geographic regions cessation of root growth of conifers in the late spring or early summer limits water uptake during the summer to the upper soil regions, which are rapidly depleted of moisture. Mycorrhizal colonization can be viewed as ' drought avoidance' in that dry soil conditions are avoided spatially through hyphal penetration of deeper zones (Parke et al. 1983).

The effects of water stress on plant growth include loss of turgor resulting in stomatal closure, reduction of photosynthetic surface, decreased biomass, reduction in photosynthetic rate and diversion of photosynthate to roots at the expense of shoot growth (Kramer, 1969). In the southeast USA, dry spring and early summer conditions can result in significant longleaf mortality upon outplanting. Results from this study, Theodorou and Bowen (1970) and Parke et al. (1983) strongly indicate that Rhizopogon can help plants to tolerate and recover from soil water deficits that can aid seedling establishment. Differences in host response to inoculation with various ectomycorrhizal fungi suggest that drought tolerance should be considered one of the more important criteria for selection of Rhizopogon species suitable for nursery inoculation of longleaf pine seedlings.

REFERENCES Amaranthus, M.P. and D.A.Perry. 1987. Effect of soil transfer on ectomycorrhiza formation and the survival and growth of conifer seedlings in disturbed forest sites. Can J For Res. 17:944-950. Amaranthus, M.P. and D.A. Perry. 1989a. Rapid root tip and mycorrhiza formation and increased survival of Douglas-fir seedlings after soil transfer. New Forests. 3:77-82. Amaranthus, M.P., J.M. Trappe, and D.A. Arthur. 1994. Hypogeous fungal production in mature Douglas- fir forest fragments and surrounding plantations and its relation to coarse woody debris and animal mycophagy. Can J For Res. 24:2157-2165. Amaranthus, M.P., D. Dumroese, A. Harvey, and L. Bednar. 1996. The effects of compaction and organic matter removal on seedling root and mycorrhizal tip production and diversity. PNW-RP 494. 12 pp. Agerer, R., W.R. Muller, and G. Bahnweg . 1996. Ectomycorrhizae of Rhizopogon subcaerulescens on Tsuga heterophylla. Nova Hedwigia. 63:397-415. Boyer, J.S. 1976. Water deficits and photosynthesis. In: vol. IV, pp. 153-190. T.T. Kozlowski (ed.). Water Deficits and Plant Growth, Academic Press, New York and London. Brownlee, C.J., A. Duddridge, A. Malibari, and D.J. Read. 1983. The structure and function of mycelial systems of ectomycorrhizal roots with special reference to their role in forming inter-plant connections and providing pathways for assimilate and water transport. Plant Soil 71:433443.

79 Dosskey, M., L. Boersma, and R.G. Linderman. 1990. Role for photosynthate demand by ectomycorrhizas in the response of Douglas-fir seedings to drying soil. New Phytol. 117:327-324. Duddridge, J.A., A. Malibari, and D.J. Read. 1980. Structure and function of mycorrhizal rhizomorphs with special reference to their role in water transport. Nature. 287:834-836. Dumroese, D., A.E. Harvey, M.F. Jurgensen, and M.P. Amaranthus. 1998. Impacts of soil compacton and tree stump removal on soil properties and outplanted seedlings in northern Idaho, USA. Can J Soil Sci. 8:77, 160-172. Johnsgard, G.A. 1963. Temperature and the water balance for Oregon weather stations. Oregon State University Agricultural Experiment Station Special Report #150. Kramer, P J. 1969. Plant and Soil Water Relationships: a modern synthesis. McGraw-Hill, San Francisco. Massicotte, H.B., R. Molina, D.L. Luoma, and J.E. Smith. 1994. Biology of the ectomycorrhizal genus Rhizopogon II. Patterns of host-fungus specificity following spore inoculation of diverse hosts grown in mono- and dual cultures. New Phytol. 126:677-690. Mexal, J.G. and C.P.P. Reid. 1973. The growth of selected mycorrhizal fungi in response to induced water stress. Can J of Botany, 51. 1579-1588. Mikola, P. 1970. Mycorrhizal inoculation in afforestation. International Review of For. Res. 3:123-196. Molina, R. and J.M. Trappe. 1994. Biology of the ectomycorrhizal genus Rhizopogon I. Host associations, host-specificity and pure culture syntheses. New Phytol. 125:653-675. Parke, J.L., R.G. Linderman, and C.H. Black. 1983. The role of ectomycorrhizas in drought tolerance of Douglas-fir seedlings. New Phytol. 95:83-95. Peterson, T.D. and M. Newton. 1985. Growth of Douglas-fir following control of snowbrush and herbaceous vegetation in Oregon. Down to Earth 41:21-25. Preest, D.S. 1977. Long-term growth response of Douglas-fir to weed control. New Zealand J of For Sci. 73:329-332. Read, D.J. and R. Boyd. 1986. Water relations of mycorrhizal fungi and their host plants. In: pp. 287-303. P.G. Ayres and L. Boddy (eds.). Water, fungi, and plants. Cambridge University Press, Cambridge. Reid, C.P.P. 1979. Mycorrhizae and water stress. In: pp. 392-408. A. Riedacker and J. Gagnaire (eds). Root physiology and symbiosis. IUFRO Symposium Proceedings. Nancy, France. Sands, R. and C. Theodorou. 1978. Water uptake by mycorrhizal roots of radiata pine seedlings. Aust J Plant Physiol. 5:301-309. Simard, S.W., D.A. Perry, M.D. Jones, D.D. Myrold, D.M. Durall, and R. Molina. 1997. Net transfer of carbon between ectomycorrhizal tree species in the field. Nature. 388:579-582. Steel, R.G. and J.H. Torrie. 1980. Principles and procedures of statistics. McGraw-Hill Book Co., New York. Theodorou, C. and G.D. Bowen. 1970. Mycorrhizal responses of radiata pine in experiments with different fungi. Aust For. 34:183-191. Theodorou, C. 1978. Soil moisture and the mycorrhizal association of Pinus radiata D Don. Soil Biol Biochem. 10:33-37. Trappe, J.M. and R.D. Fogel. 1977. Ecosystematic functions of mycorrhizae. In: pp. 205-214. J.K. Marshall (ed.). The Belowground Ecosystem: a Synthesis of Plant-Associated Processes, Colorado State University, Fort Collins. Wright, E. and R.F. Tarrant. 1958. Occurence of mycorrhizae after logging and slash burning in the Douglas-fir type. USDA Forest Service Research Note PNW-160. Portland, Oregon.

80 Fig 1. Longleaf pine seedling moisture stress at various soil moistures with and without Rhizopogon mycorrhizal inoculation 22

20 Rhizopogon inoculated 18 Control

16

14

12

10

Plant moisture stress 8

6

4

2

0 0 5 10 15 20 25 30 35 40

Soil Moisture %

Fig 2. Longleaf pine seedling moisture stress at various soil pressure levels (bars) with and without Rhizopogon mycorrhizal inoculation

22

20 Rhizopogon inoculated 18 Control

16

14

12

10

Plant moisture stress 8

6

4

2

0 .01 .1 1 10 100

Soil pressure (bars)

81 JABBERWOCKY: AN INDIVIDUAL TREE, SPATIALLY EXPLICIT, BIOLOGICALLY BASED GROWTH MODEL FOR LONGLEAF PINE AT EGLIN AIR FORCE BASE, FLORIDA

Matt Anderson (USDA Forest Service, Southern Research Station, Auburn, AL) Greg Somers ((School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849) Rick Smith (USDA Forest Service, Southern Research Station, Auburn, AL) Mickey Freeland (School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849)

ABSTRACT: An individual tree growth model called Jabberwocky is being developed by the USDA Forest Service Southern Research Station and the Auburn University School of Forestry and Wildlife Sciences. The model will serve as a decision support tool for land managers at Eglin Air Force Base in northwest Florida. Spatial detail and biologically based growth drivers enable the simulation of stand growth as an aggregate of tree growth. This will potentially simulate the growth under any conditions in which the species will grow, enabling the model to be applicable throughout the range of longleaf pine. The component-based design permits easy enhancement of the model as new data becomes available and existing relationships between various tree parts and between the tree and its growing environment are better understood. A database has been developed containing measurements made from 106 trees from a range of sites at Eglin AFB. Among the characteristics measured are the spatial location of each foliar point within the tree’s crown, annual basal area growth throughout the bole, surface area and nutrient content of needle samples throughout the crown, density of wood samples taken throughout the bole, stem taper, spatial relationship between the subject tree and competitors, and a variety of site soil characteristics. At present work is ongoing on checking the quality of the data and quantifying relationships among tree components.

INTRODUCTION A new tool is being developed to support land managers at Eglin Air Force Base in northwestern Florida. Jabberwocky is the working name given to a computer simulation model being developed by a team of researchers from the USDA Forest Service Southern Research Station and the Auburn University School of Forestry and Wildlife Sciences.

Eglin Air Force Base is comprised of over 463,000 acres of land. Most of this area is forested and serves as a buffer area around the ranges where the Air Force conducts tests of weapon guidance and tracking systems. Weapon testing helps maintain an environment of frequent low intensity fires throughout the year. This is similar to the conditions that prevailed in the area prior to settlement. Foremost among the tree species represented on the landbase at Eglin are longleaf pine (Pinus palustris Mill.), sand pine (Pinus clausa (Chapm.) Vasey.), and various scrub oaks (Quercus spp.). The soils in which these forests grow are almost all excessively well-drained sands so that even though the area receives more rainfall than any other part of Florida the trees are subject to drought-like conditions for most of the summer.

In addition to frequent fires, the location of Eglin AFB on the Gulf Coasts means that the disturbance regime influencing forest development includes periodic hurricanes. Trees blown down by Hurricane Opal in October of 1995 can be seen throughout the base.

This combination of species mixture and environmental conditions provide a habitat that is favorable to longleaf pine. The land across the base is primarily composed of all-aged stands of longleaf pine. Frequent disturbances maintain the stands in an open character and low moisture holding capacities of the soils and limited nutrients result in very slow growth.

Longleaf pine has not been subject to intensive study in the way other pine species have. Further, what little is known about longleaf pine forest ecosystems is not applicable to the forest environment at Eglin. Land managers at Eglin have few tools with which to evaluate management options, yet are responsible for the stewardship of one of the most diverse ecosystems outside of the tropics and habitat to a number of threatened and endangered species. To alleviate this shortcoming, work has begun on developing Jabberwocky, a spatially explicit, biologically based individual tree model for longleaf pine.

82 The primary purpose of Jabberwocky is to provide short-term projections of the growth and development of the longleaf pine element of forest stands at Eglin. A secondary purpose is to serve as the foundation for an ecosystem model for longleaf pine applicable to environmental conditions encountered throughout the species’ range. Smith and Somers (1998) have previously described the scope and philosophy of the model.

DATABASE CONSIDERATIONS Jabberwocky is a biologically based growth model. In the model, the quantity and distribution of photosynthetic material within an individual tree’s crown drive its growth. Water and nutrient availability and the immediate physical environment of the tree modify potential growth. The biological basis of Jabberwocky necessitates that it be scaled to the individual tree. To model the growth of whole stands of longleaf pine, information about the growth of different individual trees is combined. This requires that the ways in which the trees interact and affect each other’s growth and development be investigated.

Toward this end a number of measurement plots have been installed across the base representing the range of conditions favoring longleaf pine. On each of these plots, the location of each tree has been identified. Among trees selected for detailed measurement the location of competing trees has also been identified and the area of crown overlap between the subject tree and each of its competitors carefully mapped. To measure the tree’s photosynthetic capacity, extensive measurements of the tree’s crown are being made. Points of growth cessation over the three years prior to sampling are being identified and their three- dimensional coordinates recorded. The location and year of mortality of dead branches is also being recorded. Changes in the crown over the past three years are being determined from this data.

A subsample is being made on each tree such that every fourth branch is subjected to more intensive measurements. On these branches, measurements are made of the display of the foliage present. In addition, foliage samples are collected at one-meter intervals along the branch and measured to determine specific leaf area and nitrogen and carbon content. These measurements are being used to indicate the amount and quality of photosynthetic material on the subject tree and how it is distributed within the crown.

Spatially explicit measurements of the distribution of foliage throughout the crown and of the points of connection between each branch and the tree’s bole provide information on factors influencing the subject tree’s secondary growth. Annual basal area increment is being measured at one-meter intervals along the bole and at the base of each branch from which foliar samples are being collected. Wood density is also being measured throughout the bole.

Soil samples have been collected from each of the plots. Various physical properties of the soil such as hydraulic conductivity and moisture holding capacity have been measured. Organic matter content, cationic exchange capacities, pH, and other values have also been measured.

The effect of weather differences on site quality will be examined in detail. A gradient in rainfall amount exists across the base from east to west and from the coast inland. Weather stations located throughout the base will provide a variety of data to examine the effect of local climate on tree and stand development.

PROGRESS Preliminary work for the project saw the establishment of 107 plots on 25 different stands across the base. On these plots the location of the trees was mapped and various measurements of size taken. These measurements provided preliminary data by which the sampling frame of candidate trees for more intensive measurement was formed.

Stands across the base were divided among three classes based on site index. Ten size classes of tree were identified on each site class based on diameter and crown ratio. Intensive measurements of crown architecture, annual ring area, and wood density were carried out on trees falling within these size class limits. A second set of size classes based on diameter, crown ratio, and height was formed with forty size classes representing each site type. Less detailed measurements were made on trees within these size classes.

83 Soil samples were collected from each of the plot from which tree sample material was collected. Composite samples were collected for analysis of nutrient content at each soil horizon and organic matter at the surface. Four soil cores from each soil horizon were extracted for analysis of physical properties.

Measurements have been stored in a relational database created with Microsoft Access. Because the model is data-intensive, checking for and correcting errors has been a major undertaking. A variety of approaches have been developed to track down possible errors and to identify their source.

Preliminary modeling of relationships between the various tree components is ongoing. An area of particular interest is elucidating the relationship between sapwood area and foliage area to develop a predictor of foliage quantity based on sapwood area in accordance with the Pipe Model Theory (Shinozaki et al. 1964). Development of an accurate predictor of foliage quantity would allow for a substantial increase in the number of trees sampled as it would eliminate the costliest measurement procedures in terms of time and labor required.

Investigations of stem form have begun. Preliminary efforts have been aimed at utilizing this data set in fitting existing stem profile equations for longleaf pine (Farrar 1987). As work progresses a variety of alternate approaches of modeling stem taper will be investigated.

Other future investigations will examine the way crown shape affects light environment at the forest floor, the relationship between crown shape and wood production in different directions, and more traditional growth and yield studies.

A prototype of the model is currently being developed and coded in C++. A graphical user interface capable of rendering a graphical image of a given stand and producing reports on pre- and post-projection stand conditions is currently being tested.

LITERATURE CITED Farrar, R. M. Jr. 1987. Stem-Profile functions for predicting multiple-product volumes in natural longleaf pines. Southern Journal of Applied Forestry. 11(3):161-167 Shinozaki, K., Yoda, K., Hozumi, K., and Kira, T. 1964. A quantitative analysis of plant form: the pipe model theory. I. Basic analyses. Japanese Journal of Ecology. 14(3):97-105. Smith, R., and Somers, G. L. 1998. First steps in developing Jabberwocky: a three dimensional, biologically based individual tree model for longleaf pine. In T. A. Waldrop, ed. Proceedings of the ninth biennial Southern Silvicultural Research Conference, February 25-27, 1997, Clemson, South Carolina. General Technical Report SRS-20, USDA Forest Service, Southern Research Station, Asheville, North Carolina. pp. 107-112.

84 ARE WE KILLING TOO MANY LARGE, OLD TREES: RESTORATION EFFORTS IN AN OLD- GROWTH LONGLEAF PINE STAND

Chadwick Avery (School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849) John S. Kush (School of Forestry & Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849)

ABSTRACT: Ecological restoration reliant on the use of prescribed fire has been underway for six years in a 60-acre uncut, old-growth longleaf pine stand in Flomaton, AL. This stand had undergone > 45 years of fire suppression, leading to heavy litter accumulations, dense hardwood midstory, and a dearth of herbaceous species. Baseline data were collected prior to reintroduction of fire into the stand in 1993. Measurements were conducted in 1997 and again during the 2000 summer. There is concern about the pace of restoration and the impact these efforts may be having on the larger, older trees. This poster will present the latest growth and mortality data for the Flomaton Natural Area.

INTRODUCTION Ecological restoration of forests in North America is increasingly becoming a major topic in natural resource management. Private, state, and federal land managers have recently undertaken ecological restoration in the longleaf pine forests of the southeastern USA. Restoration to this point has lacked data and an understanding of interactions during restoration. Information on reducing litter accumulations and changes in overstory structure is lacking.

One area where the impacts of ecological restoration on longleaf pine forests are being studied is the Flomaton Natural Area in Escambia County, Alabama. The Flomaton Natural Area is a 60-acre virgin longleaf pine stand that underwent > 45 years of fire suppression. Beginning in 1993, plans began for the ecological restoration of the Flomaton Natural Area utilizing prescribed fire. Prescribed fires have been set in January or April 1995, January or April 1996, and June 1997 within the study area. In addition to the three prescribed fires, a fuelwood harvest of all hardwoods occurred in April/May 1996.

METHODS The Flomaton Natural Area is located in Escambia County, Alabama. In 1997, all longleaf pine were measured for azimuth and distance, DBH of live trees and standing snags, crown height, total height and litter depth at the base of each tree. All trees were re-measured in July 2000 to determine mortality, DBH growth and changes in litter depth.

RESULTS Longleaf pine mortality following the three prescribed fires (1995, 1996, and 1997) and fuelwood removal (1996) had been substantial, especially in the lowest diameter classes). Annual mortality for all longleaf pine over the study period, 1997-2000, averaged 13.0%; and 40.0% for all sampled trees below 3.0 inches DBH. These data contrast sharply with prior observations of less than 1% annual mortality in longleaf pine forests. Reasons for this disparity are partially associated with the exclusion of small trees in these studies. However, even with the exclusion of small trees, mortality at Flomaton Natural Area is noticeably different.

While basal area remained static (78.0 to 77.5 sq. feet/acre) from 1993 to 1997, it decreased to 53.6 sq. feet/acre. Considering the high rate of mortality and an added complication of bark consumption by fire (reducing diameter of large trees inordinately more than smaller trees), this figure is remarkable. Mean DBH of all longleaf pine shifted from 9.4 to10.8 inches from 1993 to 1997, increasing to 10.9 inches in 2000. Tree density decreased from 113 trees/acre in 1993 to 93 trees/acre in 1997 and down to 63 trees/acre in 2000.

SUMMARY Ecological restoration of the Flomaton Natural Area has been underway for 6 years. Results of data collected this year and 3 years shows changes density, basal area and overstory mortality. Monitoring of further responses of the stand to these and future restoration efforts will continue.

85 Fire is an essential element in the health and maintenance of the longleaf pine ecosystem. When fire is absent from an area inhabited by longleaf pine, many processes cease to occur in its regeneration and sustainability.

ACKNOWLEDGEMENTS The authors thank Champion International Corporation (now International Corporation) and Foster Dickard for allowing the restoration process to occur in the Flomaton Natural Area. Additionally, we thank R. Sampson, B. Chamlee, G. Ward, W. Thompson, and R. Tucker for their efforts in restoration of the Flomaton Natural Area.

86 LONGLEAF PINE GROWING IN THE SAN AUGUSTINE AND SABINE COUNTIES OF TEXAS

Clay Bales (Texas Forest Service, Rt. 5 Box 1100, San Augustine, TX 75972)

ABSTRACT: Weary travelers from many years ago entered what is today known as Texas along the historic El Camino Real otherwise known as the San Antonio Road. Although the forestland of East Texas has changed a great deal since settlement, many historical books depict this region as longleaf pine savanna across rolling hills intermingled with rich bayous. The first two Texas counties along the El Camino Real are Sabine and San Augustine Counties. Today there is less longleaf pine growing in these counties, however, efforts by private landowners, forest industry, and the US Forest Service have maintained areas in longleaf pine for economic and ecological reasons. The rise of saw-timber and pole prices, along with the decline in cattle prices, has shown an increase in longleaf pine planting. With many absentee landowners in the region, the wildlife value of longleaf pine plantations is an added benefit for a weekend retreat. The Texas Forest Service recognize that more and more landowners are becoming interested in growing longleaf pine, and they serve as a source of information for anyone interested in reforesting or managing their land.

87 ESTABLISHMENT AND MANAGEMENT OF LONGLEAF PINE (PINUS PALUSTRIS MILL.) SEED PRODUCTION AREAS

Jill Barbour (U. S. Forest Service National Tree Seed Laboratory, Macon, GA)

ABSTRACT: This paper is a how-to primer on everything you need to know about collecting longleaf pine cones in the woods. The primer is especially helpful to anyone wanting to improve their cone collection procedures. It is a compilation of research on longleaf pine cone production with practical information needed to be successful at cone collection. The paper begins with how to locate cone-producing trees concluding with shaking trees and transporting the cones out of the woods.

Stand characteristics that result in maximum cone production are specified. Flower morphology, pollination, and the associated predators that feed on reproductive structures are briefly mentioned. The cone collector is alerted to the fact that an entire crop can be lost to conelet abortion. Thinning, fertilization, and irrigation are compared to their effect on cone stimulation. Previously unpublished charts, by Dr. William Boyer, are provided for converting counted seed and cones to seed per bushel and cones per bushel. Determining cone maturation by measuring cone specific gravity is explained with instructions for building a cylinder to measure specific gravity in the woods. The proper operation of tree shakers and brand names available are discussed. Various modes of collecting fallen cones and the equipment needed are presented.

There is a growing interest in planting longleaf pine (Pinus palustris Mill.) in the south. The demand for seedlings is increasing rapidly, thus fueling the demand for more seed. The periodicity of cone crops creates a shortage of seed; therefore, more seed production areas are needed to meet the demand for seed.

To locate longleaf pine stands look in the U. S. Forest Service Resource Bulletin SRS-9 called The Longleaf Pine Forest: Trends and Current Conditions written by Kenneth Outcalt and Raymond Sheffield(1996). State district foresters or county foresters can also be contacted for the locations of longleaf pine stands in their area. The county courthouse has the landowners’ identity and coordinates.

Actively managed longleaf pine stands that are easily accessible are the best candidates for seed production areas. Cone production is highly correlated with previous fruitfulness in loblolly and longleaf pine. For maximum cone production select stands with 15 inch diameter trees that are 50 to 60 years old. The best basal area for cone stimulation is between 30 to 40 square feet which is 25 to 50 trees per acre. Above 40 square feet of basal area the number of cones per tree drops precipitously.

If you need to reduce the number of trees per acre save the best cone bearing trees. Large longleaf pine trees are favored by thinning from below which provides the extra space for large crown development that is necessary for abundant seed production. Cone crops do not benefit fully until at least 3 years after thinning.

Longleaf pine does not produce large cone crops annually. Good cone crops occur every 5 to 7 years. Heavy seed crops occur over much of the longleaf pine range once in 8 to 10 years. Average annual cone production from 1966 to 1995 at 9 locations was 21.2 cones per tree. A 5-year moving average for cone production shows an increase from < 10 cones per tree in 1970 to >40 cones per tree in 1994. From 1987 to 2000, longleaf pine cone production has tripled.

The flower crop remains fairly consistent from year to year, but lately more flowers are surviving to become cones. The flowers counts per tree averaged 45.5 and yielded 40.5 cones per tree from 1986 to 1995 at all monitored sites. At five coastal plain locations, the average percent increase in flower counts from 1984 to 1995 was 75.1, while the increase in cone production was 337.6%.

Seed production can vary from 50 seeds/cone in a good year, 35 seeds/cone in an average year, and 15 seeds/cone in a poor year.

Periodic prescribed burning prior to cone collection is necessary to reduce the understory of mid sized hardwoods for equipment accessibility and so the collectors can find the cones after they have been shaken off the trees. Permits are required to prescribe burn forest stands and they are obtained through the state

88 forestry departments. Mechanical preparation or herbicides are required when fire cannot be utilized. On the U. S. Forest Service, Escambia Experiment Forest, the cutover longleaf pine stands lost one tree per 5 acres per year in a 15-year period.

Flower primordia are formed in the summer at the tip of the branches next to the terminal and lateral buds. The female flowers are noticeable the following winter (January-March). Male flowers contain the pollen for fertilization. When the male flowers ripen (mid February to March), they will release the pollen. It takes a year for the pollen to travel to the two ovules, lodged at the bottom of each scale, to complete fertilization and form the seed. The following summer the conelets will elongate to form cones before ripening in the fall. The process takes about 20 months.

It is best to count the flowers and conelets from the ground with binoculars during the naked period (before the needles appear and the white bud is visible). When counts are made during this naked period, a conservative estimate of all the cones on the tree can be had by multiplying the binocular count by 1.5.

Counting the seed in cones is necessary to make sure the cones have enough seed to make it profitable to collect. With a sharp knife cut the cone lengthwise and count the number of full seed on the exposed face. To make the seed count easier wait a few minutes until the surrounding tissue turns brown. A delay of more than 10 minutes may brown the seed too much making counting difficult.

Conelet abortion is the process in which longleaf pine drops its conelets within the first spring after pollination. This “physiological drop”, not due to cone insects or pollination, appears to be the result of ethylene production by the foliage and expanding shoots. Conelet abortion can be as high as 90%. Cytex (CX) is an aqueous extract of marine algae and an economical source of cytokinins. It can be applied operationally with a mist blower and has shown that it can reduce conelet abortion.

The most prevalent insects that attack longleaf pine cones are Dioryctria amatella Hulst and Dioryctria abietella (D. & S.), coneworms; Laspeyresia ingens Heinrich, seedworm; and Leptoglossus corculus [Say], seedbug. Cone rust, Cronartium strobilinum, can cause heavy losses of longleaf pine conelets. Pitch canker, Fusarium moniliforme var. subglutinans, is known to infect cones and seed. The fungus is located on the seedcoat and inside the seed. The fungus, Lasiodiplodia theobromae, infects seed after the cones fall on the ground. Cones left on the ground overnight have a much higher chance of being infected with the fungus; thereby reducing seed germination.

Before you can estimate maturation you have to collect cones still attached to the tree. Do not assume the maturation date will be the same every year because weather conditions affect ripeness. Under drought conditions the date may be advanced 7 to 10 days or conversely wet weather can push back maturation 7 to 10 days. Longleaf pine cone maturation usually falls between October 1st to October 20th. The natural opening of the cones on the tree occur anywhere between October 20th to November 10th.

Cones must be mature before collection because seed yields will be greatly reduced and immature seed does not germinate or store well. Specific gravity is the measurement used to determine ripeness for cones. Longleaf pines cones are mature when the specific gravity is 0.89 or lower. Do not collect cones until 19 out of 20 cones have a specific gravity below 0.89. Float cones within 10 minutes of collection. Cones open and shed seed when the specific gravity reaches 0.70.

Greater cone stimulation with fertilization occurs in years with abundant spring and early summer rains. Average flowering of fertilized trees at the high rate in March or May was positively correlated with March- through-July rainfall in the previous year. In areas where fertilizer applications increased cone production the soils were deep sands with low nutrient holding capacity.

There are several methods of harvesting pine cones and they range in price from inexpensive to very expensive. The cheapest harvesting method, but not the most efficient, is to flail the cones with a long pole that has a hook at one end. Cone buying stations can be strategically located at individuals’ homes to pick up cones. Crates are placed at the buying stations and picked up often. Cone collectors can follow behind

89 logging operations where tree tops have been cut and left in the woods. The logging operation will have to coincide with cone maturity to make it worthwhile to collect the cones.

For mature pines a self-propelled lift can be used that elevates the cone harvester at least 60 feet. There are self propelled lifts with wire buckets at the end of telescoping arms with controls in the bucket so the operator can move the equipment while in the air. Their biggest drawback is the weight of the equipment. They do not have enough power to operate on uneven or boggy terrain. A farm tractor is needed on site to pull one out if it gets stuck.

Presently, tree shakers are the most commonly used piece of equipment to get cones out of trees. Shaking trees is the most efficient way to collect cones from large longleaf pine trees because it is the fastest method. Shakers can be self propelled or tractor mounted on the PTO. The self-propelled models have 3 wheels instead of 4 wheels. A 9 to 12 foot long metal boom, which is attached to the machine body, has a shaking head at one end. The shaker head has a scissor action as it opens and closes. There are two round rubber pillow pads at each side of the opening, which squeezes the tree trunk. The pillow pads are contained in rubber slip slings. In addition, rubber flaps cover the slip slings to allow for movement. The shaker cab usually only has room for one person.

Lubricants are used on the shaker head to eliminate debarking trees. Peanut oil is commonly used as a lubricant in the south because it is inexpensive and it can stand up to the heat. Other lubricants are grease, crisco oil, silicone spray, and graphite spray. The lubricant is liberally applied to the outside of the pillow pad, the outside of the slip sling, and the flap that covers the slip sling. The lubricant will need to be reapplied often by the operator.

Shakers have different shake patterns. There is a pecan shake pattern, a walnut pattern, and an almond pattern. The best shake pattern for pines is the walnut pattern because it has a slower speed shake than the pecan pattern and the weight of the head is closer to the tree.

The 3 wheel self propelled shakers can easily get stuck in uneven terrain. Implement tires are on the shaker when purchased; so tractor tires are recommended for shakers used in the woods because they have better traction and it increases the machine’s clearance to run over stumps. In addition, it is recommended that a diesel fuel machine be used in the woods instead of a gasoline fuel machine because gasoline engines overheat.

The shaker operator is the most important element in a successful shake. Use only well-trained, conscientious people to operate the shaker. It is best to use the same person every year once the operator becomes competent at shaking trees.

Don’t shake more trees than the cone pickers can collect in one day. Do not leave the cones on the ground overnight. Sometimes all the cones do not fall after the first shake, so the operator can come back to the tree later and shake again without hurting the tree. When it is late in the season make sure there is seed in the cones before shaking because the cones can open up and the seed fall out and then close back up overnight. The operator could be shaking trees with no seed.

Before cone collection begins, arrangements must be made to get equipment and people to the field. Having the necessary supplies and equipment at the site will help to increase the operation’s efficiency. An efficient operation will cost less over time by increasing the chances of getting the entire cone crop collected before they open on the trees.

The most common methods of pine cone collection is with 20 bushel wooden crates, plastic pickle boxes, or bushel bags. Wooden crates need to be stored under shelter or the wood will rot. Plastic pickle boxes can remain outdoors but they will deteriorate over time. A flatbed trailer is needed to deliver and collect the crates from the processing facility. A forklift is needed to load the crates on the trailer bed.

Green cones weigh about 33 to 40 pounds per bushel, just before opening 20 to 25 pounds per bushel. A 40- pound bag of cones would be difficult to carry out of the woods, so have a pickup truck or trailer to haul the

90 bags to the loading site. The bags or crates need to be labeled and include the date of collection, lot number, county, and state.

The cones loaded on the flatbed trailer can be transported to the seed processing facility once it is full or at the end of each day. Cones should not be left in bags or crates for more than a week because they are likely to mold, heat up, and ferment. Birds are the primary predators of longleaf pine seed with predation by mice and squirrels to be secondary.

It is vital to follow these guidelines to insure a successful cone harvest. Harvesting longleaf pine cones is more difficult than harvesting other southern pine cones. Diligence at every step of the process is required because any mistake can start the deterioration of the seed, thereby reducing germination.

91 THE PALUSTRIS EXPERIMENTAL FOREST—65 YEARS OF RESEARCH THAT CHANGED THE SOUTH

James P. Barnett (USDA Forest Service, Southern Research Station, 2500 Shreveport Highway, Pineville, LA 71360)

ABSTRACT: The Palustris Experimental Forest is a site located within the Kisatchie National Forest managed for research purposes by the Southern Research Station. For over 65 years, it has been the site of research that resulted in much of the technology used to establish and manage southern pine plantations. From pioneer researchers like Philip Wakeley to the current staff, a focus of the research program conducted on the Experimental Forest has dealt with reforestation and stand management. The results of these efforts have been instrumental in making pine forests a leading economic resource across the South.

INTRODUCTION The Palustris Experimental Forest was designated by Congress in 1935 as an area for conducting forestry research. The Forest is named Palustris in recognition of the species longleaf pine (Pinus palustris Mill.), which, prior to the widespread harvest of virgin pine forests in the early 1900’s, once occupied over 90 million acres in the American South. The Experimental Forest includes two separate tracts of public land within the Kisatchie National Forest that total about 7,500 acres. The Southern Research Station supervises and conducts a wide range of long-term and other studies on these unique and scientifically valuable lands.

The Palustris Experimental Forest was established due to the efforts of pioneer researcher Philip C. Wakeley with the goal of developing reforestation techniques for the four major southern pines. With the help of the Civilian Conservation Corps, Wakeley and other scientists grew seedlings at the Stuart Nursery near Pollock and outplanted them on the Experimental Forest to develop nursery technology and stock specifications for planting of southern pines.

Early research on the Palustris Experimental Forest included cone and seed studies that would become the basis for reforestation success throughout the South and around the world. Direct-seeding operations showcased a means by which denuded landscapes could be quickly reforested; and new seedling production techniques developed here pioneered the current capability of tree nurseries to produce over a billion seedlings per year. Woody plant-control methods developed on the Palustris demonstrated how unproductive sites could be converted to thriving pine forests; and how established forest stands could be intensively managed to enhance their economic productivity. Early research scientists on this premier Experimental Forest began a long tradition of multi-resource management, and their work has contributed immeasureably to sustainable resource management.

LOCATION AND EXAMPLES OF RESEARCH The J.K. Johnson Tract, located 18 miles southwest of Alexandria, Louisiana, on highway LA-488, is named in honor of one of the first industrial foresters to reforest southern pines. He worked for the Great Southern Lumber Company in southeastern Louisiana. This 2,500-acre site is home to long-term studies such as a longleaf pine planting spacing, prescribed burning, pruning, and thinning regime study that is now about 65 years old. Some areas of the Johnson Tract are used for shorter-term studies that allow scientists to evaluate seedling physiology. In addition, innovative research is now underway to evaluate the effects of global climate change on forest productivity and to devise management strategies to reduce such effects. Using very intensive measurements to tree and stand morphology and physiology, studies like this involve ongoing, cooperative efforts with a full range of partners. To facilitate those efforts, the Southern Research Station has made lodging available for cooperating scientists, graduate students, and summer interns. Research conducted on the Johnson Tract is now benefiting from some of the most sophisticated equipment available to plant scientists.

The Longleaf Tract, about 35 miles south of Alexandria, Louisiana, off highway US-165 near Glenmora, was added to the Experimental Forest in the late 1940’s. It has been the site of some of the most intensive multi- resource research in the South. Since the 1940’s, the interactions of cattle grazing, wildlife management and timber production have been evaluated. This research was critical to the reforestation efforts following World War II. During the War, the vast cutover longleaf pine forests become grasslands where herds of

92 cattle and hogs roamed freely. Success of reforestation efforts required that the interactions among cattle, hogs, and trees be understood. This effort has provided the information necessary to allow integration of grazing, wildlife habitat, and forest productivity. Current research emphases include evaluations of effects of forest management practices on long-term soil productivity. With burgeoning human demands for forest products and amenities, scientists are using the Longleaf Tract to evaluate in depth the effects of timber harvesting, prescribed fire, site preparation, and pine straw utilization on soil structure, nutrition, and chemistry; the ecology of soil microorganisms; soil-plant moisture relationships; and plant productivity.

CLIMATE The climate is typical of the southern Coastal Plain. Annual precipitation averages 57.7 inches with fairly equal monthly distribution. During the Winter and Spring, the average rainfall is 29.7 inches compared to 28.2 in the Summer and Fall. October is usually the driest month. Temperatures average 72 °F with minimums of 23 °F and maximums of 96 °F.

DATA BASES Monthly rainfall and temperature data have been collected since 1950. These data are being made available in machine-readable format. Hourly maximum/minimum air temperatures, soil temperature, solar radiation, wind speed, and precipitation have been collected using automated weather stations since 1986.

Numerous long-term (30 to 60 years) growth data sets have been collected for longleaf, loblolly, and slash pine. These data are the basis of growth and yield prediction systems that have been developed for these species. Other studies quantifying intensive soil and tree physiology measurements have been underway for about 10 years.

FACILITIES Each of the Tracts has facilities that include offices, laboratories, warehouses, shops, and equipment sheds. In addition, the Longleaf Tract has a residence and the Johnson Tract has lodging available for cooperating scientists, graduate students, and summer interns.

BIBLIOGRAPHY OF IMPORTANT PUBLICATIONS Research conducted on the Palustris Experimental Forest has led the development of reforestation technology in the South. Because the harvesting methods of the virgin forests were efficient by the time they moved into the West Gulf Region, emphasis on reforestation techniques were essential to reclaim the cutover land. Examples of the nearly 1,500 publications that document the development of this technology are listed below:

REFERENCES Wakeley, Philip C. 1954. Planting of the southern pines. Agriculture Monograph No. 18. Washington, DC: U.S. Department of Agriculture. 233 p. Campbell, R.S.; Peevy, F.A. 1950. Poisoning certain undesirable hardwoods for forest and range improvements. The American Midland Naturalist 44(2): 495-505. Cassady, J.T. 1953. Herbage production on bluestem range in central Louisiana. Jour. Range Management 6: 38-43. Derr, H.J.; Mann, W.F., Jr. 1953. Cost of pruning longleaf pine. Jour. Forestry 51: 579-580. Derr, H.J.; Cossitt, F.M. 1955. Longleaf pine direct seeding. Jour. Forestry 53: 243, 246. Mann, W.F., Jr.; Whitaker, L.B. 1955. Effects of prescribe burning 4-year-old planted slash pine. Fire Control Notes 16(3). Derr, H.J. 1957. Effects of site treatment, fertilization, and brownspot control on planted longleaf pine. Jour. Forestry 55: 364-367. McLemore, B.F. 1959. Cone maturity affects germination of longleaf pine seed. Jour. Forestry 57: 648- 650. McLemore, B.F.; Czabator, F.J. 1961. Length of stratification and germination of loblolly pine seeds. Jour. Forestry 59: 267-269. Echols, H.W. 1966. Texas leaf-cutting ant controlled with pelleted mirex bait. Jour. Econ. Entomology 59: 628-631.

93 Derr, H.J. 1966. Longleaf x slash hybirds at age 7: survival, growth, and disease susceptibility. Jour. Forestry 64: 236-239. Shoulders, E. 1967. Fertilizer application, inherent fruitfulness, and rainfall affect flowering of longleaf pine. Forest Science 13: 376-383. Derr, H.J.: Mann, W.F., Jr. 1971. Direct-seeding pines in the South. Agricultural Handbook 391. Washington, DC: U.S. Department of Agriculture, Forest Service. 68 p. Mann, W.F., Jr.: Dell, T.R. 1971. Yields of 17-year-old loblolly pine planted on a cutover site at various spacings. Research Paper SO-70. USDA Forest Service, Southern Forest Experiment Station, New Orleans, LA. Barnett, J.P.; Brissette, J.C. 1986. Producing southern pine seedlings in containers. Gen. Tech. Rep. SO-59. USDA Forest Service, Southern Forest Experiment Station, New Orleans, LA. 71 p.

CONCLUSIONS Great strides have been made in the last 60 to 70 years to develop the capability to restore the productivity of our southern forest lands. These forests are now the mainstay of the economies of all the southern states. The research conducted on the Palustris Experimental Forest has been instrumental in providing the knowledge that is the basis of understanding of how to sustain the productivity of our southern pine forests.

94 INTERACTIONS OF FIRE, INSECTS AND ROOT PATHOGENS IN LONGLEAF PINE

T.A. Bauman (LSU School of Forestry, Wildlife and Fisheries, Baton Rouge, LA) L.G. Eckhardt (LSU Dept. Of Plant Pathology, Baton Rouge, LA) R.A. Goyer (LSU School of Forestry, Wildlife and Fisheries, Baton Rouge, LA) K.D. Klepzig (USDA Forest Service, Southern Research Station, Pineville, LA)

ABSTRACT: Longleaf pine is considered to be relatively tolerant of, or resistant to, most insects and diseases that affect southern pines. Increased mortality, however, is being observed in some 30-45 year old longleaf pine stands within two years after burning. Knowledge of attacking insects and the root-infecting fungi that are associated with prescribed or wildfires would provide important information relevant to understanding their role(s) with fires. This work-in-progress describes a cooperative study investigating the interaction of fire intensity and burning regime with the occurrence and abundance of beetles and weevils that can be primary or secondary invaders. This study describes, also, the role of these insects as vectors of root pathogens, primarily Leptographium spp.

Potential insect pests and vectors were collected from flight interception traps and pitfall traps baited with ethanol and pine turpentine. Root samples were collected using a modified two-root excavation method. Scolytidae and collected were counted, identified and used to isolate Leptographium spp. Potential vectors and root samples were plated on selective media (2% malt extract agar, MEA, containing cycloheximide and streptomycin sulfate). Rates of insect occurrence and fungal infection will be reported.

------Longleaf pine is considered to be relatively tolerant of, or resistant to, most insects and diseases that affect southern pines. However, the exclusion of fire from longleaf pine stands has encouraged competition from other species, such as loblolly pine and various hardwoods. To restore longleaf pine sites to their earlier composition, several propositions involving the use of prescribed fire have been vigorously pursued.

Fire, either ‘natural’ or prescribed, is useful in removing undesirable species as well as stimulating longleaf reproduction by reducing the incidence of brown spot needle blight, and thus, allowing release of seedlings from the grass stage. With the re-introduction of fire into longleaf pine ecosystems has come an observed increase in tree mortality in 30-45 year-old longleaf stands within two years of burning. Although the factors involved in this increased decline are not well understood, several biological and tree physiological factors such as decreased resistance to insects and diseases are suspected.

In this study, we are investigating the effects of fire on tree health as influenced by the interaction of fire intensity, burning regime (timing), and increased susceptibility to insects and vectored pathogens. In preliminary experiments conducted in the Kisatchie National Forest in central Louisiana, we have separated longleaf pines into categories of crown scorch, bole char, and timing of fires, and are monitoring both the abundance and occurrence of several known and suspected insect pests and pathogens of disease that they vector. To date, three prescribed burns- one dormant season burn and two growing season burns– have been conducted. Additionally, one wildfire in a wilderness area that occurred in July 2000 is presently under comparative study.

When to use fire to manage longleaf stands is becoming an increasingly important question. Growing season burns may become intense and damage trees, making them susceptible to insects and disease. However, some dormant season burns may not be hot enough to remove unwanted understory, and, in some cases, may increase the re-sprouting of understory following the fire. Alternatively, dormant season burns also can be slow moving and may expose trees to high temperatures for a longer period of time, damaging the cambium and causing girdling. The flora associated with growing season fires is different from the flora associated with dormant season fires. Determining the intensity and timing of fire that will achieve the desired condition of the longleaf stand and not cause deleterious damage to the trees is the objective of this investigation. In an effort to quantify fire intensity and damage to tree cambium, we are using a propane flamethrower with a datalogger and thermocouples, and mechanical damage to trees to determine impacts and potential mortality. We are also using thermal paints during prescribed burns to verify the temperatures trees were subjected to.

95 In response to fire, trees produce volatiles that serve as chemical attractants to pest insects. Also, trees produce resin as a means of defense against insect invasion. Of the insects that respond to fire induced volatiles, engraver beetles (Ips sp.), pitch weevils (Pachylobius and Hylobius spp.), black turpentine beetles (Dendroctonus terebrans) and spp. are insects that may have significant ecological and physiological effects on trees after a fire. These insects generally attack trees that have reduced resistance due to stress or damage, but nearby healthy trees also may be attacked. Also, these insects cause significant damage to the host tree that may lead to mortality by mechanical damage to the tree cambium and/or introduction of fungal pathogens. These fungi are adapted to insect dispersal and are commonly associated with bark beetles, such as Ips engraver beetles, many of which infest conifers, including longleaf pine. A number of Leptographium species of fungi have been associated with pine decline and mortality, primarily as associates of root feeding bark beetles and weevils, such as the black turpentine beetle and pitch weevils, that attack living trees. The Leptographium wageneri teleomorph, Ophiostoma wageneri Harrington, is associated with Hylastes macer LeConte (Coleoptera: Scolytidae). However, the biological basis for these relationships is variable, and the relative importance of the beetles and the fungi in killing the trees is still unclear. Trees are killed by the fungi by the blockage of water transport in the outer annual rings. Changes in twenty-four hour resin flow rates are used to quantify the tree’s response to fire intensity and resistance to insect invasion. Research to date indicates that trees significantly increase resin production for several months following a fire, when compared to the unburned control. This increase may help to prevent successful attacks by insects attracted to the burn. However, preliminary observations from earlier fires indicate that resin flow may ‘crash’ if cambial injury has occurred, resulting in virtually no resistance to attacking insects.

In our Kisatchie study, we trapped insects after the March 2000 burn, and for six weeks prior to a prescribed burn in July of 2000, using both flight interception and pitfall traps. The traps were baited with ethyl alcohol and turpentine and checked weekly, and the number of potential pest insects captured, were recorded. We continued to trap insects for seven weeks after the July 2000 fire. To date, fire is shown to significantly increase the number of insects immigrating into the area. These numbers remain high for weeks after the fire and then either remain high, as in the March 2000 fire, or appear to plateau and fall to levels similar to unburned areas, as in the July 2000 fire.

Root samples were collected from the May 1999 and March 2000 fires. Root samples for the July 2000 burn currently are being collected. Trees for sampling were chosen according to scorch class: low to moderate, and severe. Leptographium sp., Graphium sp. and Ophiostoma sp. were isolated from these samples by plating the roots on CSMA selective media and currently are being speciated. Roots from trees with severe scorch (March 2000) were found to have burn damage and excessive weeping. Trees with severe scorch had in excess of 50 visible insect attacks and had a high incidence of Leptographium sp. The incidence of Graphium sp. and Ophiostoma sp. were low. Of the trees selected, the low to moderately burned trees had 0 to 20 visible insect attacks and a very low incidence of Leptographium sp. These trees did, however, have a higher incidence of Graphium sp. For the March 2000 burn, all insects determined to be ecologically important, were found to carry Leptographium sp., Graphium sp. and Ophiostoma sp.

Due to ambient conditions and fuel moisture levels, the dormant season fire in March was more intense than the growing season fire in July, and insect response was high and moderate, respectively. This may indicate a linear rate of response by insects to the intensity of fire. Other factors, such as the prolonged drought and abnormally high temperatures, may have had an effect on insect populations in the areas under study. The 1999 growing season burn demonstrated that trees with high levels (>70%) of crown scorch and intense bole char had increased mortality and associated insect and disease incidence.

This is a work in progress, and as such, we cannot draw conclusions based on these preliminary data. However, it is clear that fire both increases the abundance of potentially damaging insect species and that very high proportions of these insects are carrying fungi suspected of causing root and other debilitating diseases to longleaf pines. With the occurrence of fire, either as a means of stand management or as a ‘natural’ wildfire, the interacting roles of insects, fungi, and the fire itself, may lead to a decline in tree health, and thus, a higher incidence of stand mortality. A better understanding of the roles each plays in tree health is necessary in order to achieve the desired longleaf stand condition.

96 INFLUENCE OF HERBACEOUS VEGETATION CONTROL ON NATURALLY REGENERATED LONGLEAF PINE SEEDLINGS

Anne Carraway (School of Forestry and Wildlife Sciences, 108 M White Smith Hall, Auburn University, AL 36849) Dean Gjerstad (School of Forestry and Wildlife Sciences, 108 M White Smith Hall, Auburn University, AL 36849) Rhett Johnson (Solon Dixon Forestry Education Center, Route 7, Box 131, Andalusia, AL 36420)

ABSTRACT: Little investigation has been done involving the response of naturally regenerated longleaf pine seedlings to herbaceous, specifically grass, control although many have suggested the importance of its influence on seedling growth. This study intends to fill gaps in research and understand how naturally regenerated longleaf pine seedlings respond to varying types of vegetation control. Four treatments: control, fire, herbicide (Fusilade DX), and fire and herbicide were applied to sites in south Alabama. Data collection consisted of line intercept sampling and biomass sampling done pre-treatment, two months and one year post treatment. Fusilade DX proved to be an effective source of control for grasses resulting in, after one year post treatment, a decrease of 98% (herbicide) and 90% (fire and herbicide) of pre treatment grass cover, whereas the control plots lost only 1% of pre-treatment grass cover. Results of biomass sampling after one year show significant difference in longleaf weights between the burn treatment and the control and herbicide treatment, with the burn weights being lower. The results of the line intercept sampling also show the burn treatment to have the greatest reduction in percent cover of longleaf among all four treatments with a reduction of 74% of the pre-treatment longleaf cover. The herbicide treatment showed the greatest percent increase in cover after a year, increasing 21% above pre-treatment numbers whereas the control plots lost 44% of pre-treatment cover. Results indicate that removing the competition of grasses from naturally regenerated longleaf seedlings will yield an increase in longleaf coverage. However, with an average pre- treatment root collar diameter in this study of 7mm, the results also imply that burning, even after reducing grass cover, is highly deleterious to longleaf seedlings that have not reached the recommended root collar diameter for fire survival of 8mm.

INTRODUCTION Longleaf pine has long been known to be intolerant to competition, especially in the grass stage. However, much of what has been studied looked at the effect woody vegetation competition had on seedlings. Few studies investigated how herbaceous vegetation, mainly grasses, influences seedlings. Research has shown that grass will act as a shield for seedlings, protecting them from temperature extremes. However, when a fire runs through a stand the grass will burn at high temperatures and kill seedlings. Other studies have suggested that grass might suppress growth and influence the survival of seedlings without the influence of fire. A separate study (see other paper by Carraway) following the fate of a bumper seed crop suggest herbaceous vegetation, compromised mainly of grasses, to have more influence on seedling numbers than woody vegetation (Figures 1 and 2). These studies did not, however, investigate the response of naturally regenerated longleaf pine to release from grass competition.

Seedlings per Acre and Herbaceous Competition

25000 1997 20000 1998

15000

10000

5000 Seedlings per Acre 0 0 20 40 60 80 100 Percent Cover

Figure 1.

97 Seedlings per Acre and Woody Competition Woody and Acre per Seedlings

25000 1997 20000 1998 15000

10000

5000 Seedlings per Acre 0 20406080100 0 Percent Cover Percent

Figure 2.

This study intends to fill the gaps in research and understand what influences grass has on naturally regenerated seedlings growth and survival.

METHODS Research was conducted at the Solon Dixon Forestry Education Center located about 30 kilometers south of Andalusia, Alabama. The Center covers 2,144 hectares, including diverse forest types such as bottomland hardwood, baldcypress swamps, and longleaf pine ridges. The stands used in this study were naturally regenerated longleaf pine stands 40 to 60 years.

Five 1 hectare blocks were set up across the Center in uniform longleaf stands. Each block was divided into .25 hectare plots. Each plot was assigned a treatment. Treatments used, including a control, were fire, herbicide, and fire & herbicide. The herbicide used was Fusilade DX which kills only grasses and does not effect forbs or longleaf. The Fusilade was applied at a rate of 63.35 oz/ha and was applied multiple times. For the fire & herbicide treatment, the herbicide was sprayed before and after the burn. The burns were done during the growing season.

In each plot, four 10m transects were established for line intercept sampling. Sampling was done pre- treatment, two months post-treatment, and one year post-treatment. At sampling two 1 biomass clip plots were taken per plot. These were separated into vegetative groups (grass, forb, shrub, tree, longleaf), dried, and weighed. Root collar diameters were also measured on 10 longleaf seedlings at each transect pre- treatment.

RESULTS AND DISCUSSION Results from biomass sampling showed the weight of longleaf pine to be significantly lower for the fire treatment than the control and herbicide (Figure 3). The fire & herbicide treatment also had longleaf weights lower than than control and herbicide, though not significantly lower. This suggests seedlings which have relatively small root collar diameters as these did with an average of 7mm, that burning, even after reducing grass cover, is highly deleterious to seedlings.

Results from the line intercept method suggest that Fusilade is a successful method for controlling grasses. Those plots treated with herbicide experienced a reduction of 98% (herbicide) and 90% (fire & herbicide)

98 Biomass of Longleaf One Year Post-Treatment

30 a a 20

10 ab b Weight (g) 0 Control Fire Herbicide Fire X Herbicide Treatment

Figure 3. cover in grasses compared to pre-treatment data, whereas the control lost only 1% (Figure4). Longleaf coverage increased only in the herbicide treatment, increasing by 21%, which is more significant when compared to the control where longleaf coverage decreased by 44% compared to pre-treatment coverage (Figure 5).

Results from this project suggest the following. First, that burning before seedlings can reach a safe root collar diameter of 8mm, even if removing competing and hot burning grass, is not beneficial. Second, that removing grass competition will not increase actual longleaf biomass but will possibly allow for seedlings to spread out and take up more space. A longer-term study may detect consequences and/or benefits of allowing seedlings to spread out.

Pecent Change in Cover of Grasses Across Treatments

100

50

0

-50 Percent Change Percent -100 Control Fire Herbicide Fire X Herbicide Treatments

Figure 4.

Percent Change in Cover of Longleaf Pine Across Treatments

100 50 0 -50

Percent Change -100 Control Fire Herbicide Fire X Herbicide Treatments

Figure 5.

99 1996 BUMPER LONGLEAF PINE SEED CROP SURVIVAL STUDIES AT THE SOLON DIXON FORESTRY EDUCATION CENTER

Anne Carraway (School of Forestry and Wildlife Sciences, 108 M White Smith Hall, Auburn University, AL 36849) Dean Gjerstad (School of Forestry and Wildlife Sciences, 108 M White Smith Hall, Auburn University, AL 36849) Rhett Johnson (Solon Dixon Forestry Education Center, Route 7, Box 131, Andalusia, AL 36420)

ABSTRACT: Longleaf pine is a poor producer of seed compared to other southern pines. This species produces a seed crop adequate for regeneration only every 5-7 years. When one of these bumper seed crops occurred in Fall of 1996 in southern Alabama an opportunity arose to follow its fate and understand more about the somewhat elusive grass stage of this species. Plots were set up in thirty naturally regenerated longleaf pine stands across the Solon Dixon Forestry Education Center in Andalusia, Alabama in summer of 1997. Information collected in these plots included basal area of longleaf and all other mature tree species, vegetation cover, litter depth, and seedling counts. A census of these plots has occurred every summer since 1997. Results show those stands burned within a year before seed fall had the highest number of seedlings germinate in 1997 compared with those stands which were burned over a year before seed fall. Data collected also shows that with greater litter depth there will be lower seeding numbers. This supports the need to burn relatively close to seed fall to ensure germination. When looking at vegetation cover, the results suggest that herbaceous vegetation has a greater impact on seedling numbers than woody vegetation. In other words, seedlings are more tolerant of woody vegetation cover than they are of herbaceous vegetation cover. As of 1999, 45% of the stands had less than the desired 500 seedlings per acre to be considered successful regeneration. No previous research states what a normal and acceptable percentage is, therefore it is unknown at this time if these stands are displaying normal regeneration numbers. A census of these stands will continue to take place each summer and results will be updated.

INTRODUCTION Longleaf pine has historically been recognized as a poor seed producer relative to other southern pine species. Longleaf only produces a seed crop adequate for successful regeneration every 5-7 years. With such infrequent bumper seed crops it is imperative to ensure successful germination and subsequent establishment.. To do so requires an understanding of how site characteristics, including basal area, litter depth, and vegetative cover interact and effect longleaf seedlings.

In 1996, there was a bumper seed crop throughout the Southeast and an opportunity arose to follow the fate of the seed crop. The life history of a longleaf pine seed crop has not yet been documented to the extent of collecting data on a long term basis that includes not only information on the longleaf seedlings but the characteristics of the site as well.

METHODS This study took place at the Solon Dixon Forestry Education Center located about 18 miles south of Andalusia, Alabama. The center covers 5,300 acres, including diverse forest types such as bottomland hardwood, baldcypress swamps, and longleaf pine ridges.

In spring 1997, monitoring plots were established in 30 stands of naturally regenerated longleaf pine 40 to 60 years old. In each stand 10 points were set up in a grid fashion. Each point represented the plot center for three circular nested plots. Plots sizes were 1/10th acre, 1/100th acre, and 1/1000th acre. In the 1/10th acre plot the species and diameter at breast height (DBH) were taken for all trees in the plot with a DBH greater than 4 inches. In the 1/100th acre plot the percent cover was estimated for each of the following vegetative groups: grasses forbs, woody shrubs, and vines, as well as percent bare ground. In the 1/1000th acre plots the longleaf seedlings were counted and the litter depth was measured.

Measurements were taken during the summers of 1997, 1998, 1999, and 2000. Summer 2000 data has yet to be analyzed.

100 RESULTS AND DISCUSSION Looking at Table 1 we can see how beneficial burning within the year of seed fall is. Those stands that were burned within the year of seed fall had much higher numbers of seedlings germinate, over 61,000 per acre, than those that were burned over a year before seed fall which had no more than 13,000 seedlings per acre. Although those stands that had the highest numbers of germinated seeds also had the highest basal area, 39 ft2/acre compared to 22 ft2/acre and 33 ft2/acre, the large difference in number of germinates makes up for the confounding factor. The litter data gives more support to the time of burn issue. Figure 1 shows the number of seedlings against the depth of litter and the trend shows higher seedling numbers with lower litter depth. This reinforces the need to burn and reduce the litter layer within the year of seed fall.

Basal Area Seedlings/ ft2/acre Burn Acre

93 Growing 1996 61,641

22 Growing 1995 8,871

33 Before 1995 12,519

Table 1.

Number of Seedlings and Litter Depth

25000 20000 15000 10000 5000

Seedlngs per Acre 0 0 0.5 1 1.5 2 2.5 Litter Depth (inches)

Figure 1.

The vegetation data suggests that herbaceous vegetation has a greater influence over longleaf pine seedlings than woody vegetation does. Figure 2 shows seedlings in 1997 growing in areas where the woody vegetation comprises between 30% and 90% of the cover and in 1998 the seedlings, though reduced in number are growing within that same parameter between 40% and 75%. The herbaceous vegetation (Figure 3) suggest something different. In 1997 the seedlings were growing in areas where the herbaceous vegetation comprised between 15% and 60% of the cover. The next year, 1998, shows the seedlings having reduced numbers and only growing where the herbaceous cover comprised 0% to 25% of the cover (with the exception of one plot at 50%). What this data suggest is that longleaf pine seedlings will germinate in relatively higher percents of vegetation cover, both woody and herbaceous, but after a year when the woody vegetation appears to have little influence on seedling numbers the herbaceous vegetation seems to influence the seedlings in a detrimental way. The more herbaceous cover there is, the fewer the number of seedlings.

The 2000 data will be incorporated into that which has already been collected and the plots will continue to be censused.

101

Seedlings per Acre and Woody Competition

25000 1997 20000 1998

15000

10000

5000 Seedlings per Acre Seedlings 0 0 20406080100 Percent Cover

Figure 2.

Seedlings per Acre and Herbaceous Competition

25000 1997 20000 1998

15000

10000

5000 Seedlings per Acre 0 0 20406080100 Percent Cover

Figure 3.

102 A SYSTEMS APPROACH FOR LONGLEAF PINE ECOSYSTEM RESTORATION

Terry R. Clason (LSU AgCenter, Hill Farm Research Station, Homer, LA 71040) Donald P. Reed (LSU AgCenter, School of Forestry, Wildlife, and Fisheries, Louisiana State University, Baton Rouge, LA 70803)

ABSTRACT: Species composition, biological structure and functional dynamics that characterize longleaf pine ecosystems are influenced by soils, topography, hydrology, and geographic location. Since species composition influences both structure and dynamics of the system, establishing and maintaining the plant species matrix is the key to a successful restoration program. A restoration project was established the LSU AgCenter Lee Memorial Forest in Sheridan, Louisiana. Forest management practices that include longleaf pine reforestation methods, herbaceous weed suppression, and prescribed burning are being used to develop environmentally sound, cost-effective, conservation management systems for longleaf pine ecosystem restoration. The study area is a recently thinned 30-year mixed pine stand having a residual longleaf overstory basal area of 15 ft2/acre. Prior to harvest, 27 0.5-acre plots were established, and the understory vegetation was inventoried by composition and density for each plot. Following the harvest, longleaf pine seedlings were planted at 225 trees/acre, and nine factorial treatment plots were replicated three times within a randomized block design. Factorial treatments included three herbaceous weed suppression treatments (untreated check, spot treatment centered over each seedling, and band treatment centered on the planted row) and three prescribed burning treatments (annual growing season burn, biennial growing season burn, and triennial growing season burn). Initial understory vegetation inventory identified 160 plant species distributed among 105 genera. Ten arborescent genera accounted for 16 species, while remaining plants in the matrix were non-arborescent, forb and grass species. The initial impact of the management practices on longleaf seedling establishment, natural and planted, and changes in the understory species matrix will be reported.

INTRODUCTION Longleaf pine ecosystem restoration means restoring species composition, biological structure, and functional dynamics that characterized the longleaf system. These characteristics are influenced by soils, topography, hydrology, and geographic location and vary throughout the natural range of longleaf pine. Since species composition influences both structure and dynamics of the system, establishing and maintaining the plant species matrix is the key to a successful restoration program. A longleaf pine ecosystem restoration research project was initiated by the Louisiana Agricultural Experiment Station at the Louisiana State University, Lee Memorial Forest, in Sheridan, Louisiana. Forest management practices that include prescribed burning, competing vegetation suppression, and longleaf pine reforestation methods will be used to develop environmentally sound, cost-effective, conservation management systems for longleaf pine ecosystem restoration.

OBJECTIVES 1. Determine impact of herbaceous weed suppression on the understory plant species matrix. 2. Determine impact of herbaceous weed suppression on planted longleaf pine seedling growth and development. 3. Determine impact of burning frequency on planted and natural longleaf pine seedling growth and development.

METHODS The study area is a 22-acre mature, mixed pine stand of longleaf, loblolly, slash, and shortleaf pine having understory vegetation rich in species diversity. In 1996, each overstory pine was stem mapped by species and size, and the understory vegetation was inventoried by composition and density along transect lines placed at 33 foot-centers. Growing season burns were applied to the entire stand in 1996 and 1997. A seed tree harvest in the fall of 1998 left a residual longleaf overstory basal area of 26 ft2/acre. Containerized longleaf pine seedlings were planted in February 1999 at a density of 110 seedlings per acre.

Following the seed tree harvest, the study area was divided into three 7-acre blocks that varied by residual longleaf stocking density (Figure 1). Each block was divided into nine treatment plots and nine factorial treatments were randomly assigned within each block. Factorial treatments were comprised of three

103 herbaceous weed suppression application methods and three prescribed burning regimes. Weed suppression application methods included 1) Band method applied as a 4-foot wide swath centered over a row of planted seedlings, 2) Spot method applied in a 2.3-foot radius circle centered over each seedling, and 3) Untreated check. Prescribed burning regimes included 1) Annual growing season burns from 2000 to 2005, 2) Biennial growing season burns from 2001 to 2005, and 3) Triennial growing season burns from 2002 to 2005. These burning regime treatments will be used to determine the impact of burning frequency on natural and planted longleaf seedling growth and development, and understory plant species matrix development.

Two longleaf pine reforestation methods, natural regeneration, and planted seedlings, were established within each factorial treatment plot. Natural regeneration was evaluated in ten 2.3-foot radius circular plots randomly distributed along the understory inventory transect lines. Longleaf pine seedlings were planted at 12-foot intervals along the transect lines. Herbaceous weed suppression treatments were hand applied in the spring of 1999, using a tank mix of hexazinon (8 oz ai) and sulfometuron-methyl (2 oz ai) per acre.

Treatment impact on the ecosystem will be monitored for 8 years from 1998 to 2005. Data collection includes residual seed tree growth and seed production, natural and planted seedling survival and growth, and herbicide impact on understory plant species matrix. Initial survival rate for planted longleaf pine seedlings and changes in understory plant species matrix are being reported

RESULTS Initial understory vegetation inventory identified 160 plant species distributed within 105 genera. There were 10 arborescent genera that accounted for 16 species. The remaining plants in the species matrix were non arborescent, forb, and grass species. In addition, the understory contained loblolly, slash, longleaf, shortleaf, and spruce pine seedlings. Loblolly, slash, and longleaf were the predominate overstory pines having a combined stocking density and basal area of 115 trees/acre and 129 ft2/acre. Following the seed tree harvest, pine overstory consisted of longleaf pine having a stocking density and basal area of 31 trees/acre and 26 ft2/acre. Residual overstory varied by replication (Fig. 1).

In 1996, average number of plant species detected in each inventory plot was 23. Although the average number of plant species decreased to 11 in 2000, species diversity did not differ among herbaceous weed suppression treatments, distribution among grass, forb, and woody species remained the same, and the predominate species from 1996 were still present (Table 1). The decline in species diversity may be attributed difference in inventory dates, 1996 inventory was conducted during a wet spring and the 2000 inventory was conducted in the fall following a dry growing season. Seed tree harvesting operation may have affected understory diversity. However, harvesting pressure was minimal on replication 3, and species diversity was similar to replications 1 and 2.

Table 1. Predominate species present in the understory plant matrix in order of abundance

Grass Species Forb Species Woody Species

Andropogon sp. Aster sp. Quercus sp.

Dichanthelium sp. Lespedeza sp. Vaccinimu sp.

Chasmanthium sp. Eupatorium sp. Rubus sp.

Aristida sp. Solidago sp. Carya sp.

Paspalum sp. Tephrosia sp. Diospyros virginiana

Herbaceous weed suppression influenced planted seedling survival with survival rate averaging 69, 57, and 62 percent for the band application method, spot application method, and the untreated check treatments. Seedling survival following the first annual burn averaged 57 percent and was less than the unburned biennial and triennial regimes which averaged 69 and 61 percent, respectively.

104 CONCLUSIONS After two years, management interventions, herbaceous weed suppression combined with burning, had no detectable impact on the understory plant species matrix. Band application method for herbaceous weed suppression improved seedling survival rate, while the spot method appeared to reduce survival rates. Early introduction of fire may be detrimental to planted longleaf pine seedling survival.

Figure 1. Longleaf pine overstory stocking by replication expressed as stand density and basal area.

105 REINTRODUCING FIRE TO HARDWOOD-DOMINATED DEPRESSIONS IN A LONGLEAF PINE-WIREGRASS SAVANNA: AN ADAPTIVE MANAGEMENT APPROACH

K. L. Coffey (Joseph W. Jones Ecological Research Center, Newton, GA 31770) L. K. Kirkman (Joseph W. Jones Ecological Research Center, Newton, GA 31770) S. B. Jack (Joseph W. Jones Ecological Research Center, Newton, GA 31770) L. M. Conner (Joseph W. Jones Ecological Research Center, Newton, GA 31770)

ABSTRACT: Karst depressions are embedded within longleaf-pine wiregrass savannas of southwestern Georgia, have saturated soil conditions or standing water for short periods, and are frequently dominated by oaks. Oak invasion is exacerbated with winter season, low-intensity prescribed fire. Once established, oaks appear to alter the ground cover as well as potential fire regime. Because the flora and fauna associated with wet-mesic savannas are diverse, restoration of this habitat is a priority in conservation management for biodiversity at a landscape scale. Oak-dominance transforms these sites such that hardwood removal is necessary prior to re-introducing fire into these systems. Prior to large-scale restoration, we are using an adaptive management approach to document rates of change of flora and fauna associated with restoration management techniques. In 12 wetland depressions surrounded by fire-maintained longleaf pine-wiregrass communities at Ichauway, we have assigned one of three experimental restoration techniques: a) fuel-loaded, high-intensity fire; b) mechanical/chemical removal of hardwoods; or c) no treatment. Specific questions addressed include: 1) At what rate does vegetation composition and species richness change with reintroduction of fire? 2) Which ecotonal and wet savanna species reappear without reintroduction? 3) Which techniques of hardwood removal most rapidly promote fuel to re-introduce fire? 4) What changes in faunal (vertebrate) species are associated with changes in vegetation? 5) What influence does oak dominance have on water budget within depressions? On-going measurements include monitoring vegetation, depth and duration of inundation, fuel accumulation, and use of habitat by amphibians, birds and bats.

106 LONGLEAF PINE DYNAMICS ON A FLATWOODS SITE: A STUDY ON THE CROATAN NATIONAL FOREST

Susan Cohen (USDA Forest Service, Southern Research Station, Research Triangle Park, NC) John S. Kush (School of Forestry and Wildlife Sciences, Auburn University, AL) Kim Ludovici (USDA Forest Service, Southern Research Station, Research Triangle Park, NC)

ABSTRACT: Natural regeneration of longleaf pine is one of the most important management tools natural resource managers have at their disposal to perpetuate existing longleaf pine stands in the southern United States. Some studies indicate a tendency for longleaf to regenerate in gaps within the already open park-like stand structure. However, high variation and unpredictability in year-to-year cone production make natural regeneration problematic. Most longleaf pine gap studies have been conducted on sandhills or excessively drained sites; however, the more poorly drained flatwood and savanna sites are generally more productive and contain higher numbers of rare and endangered species. Research sites on the eastern Coastal Plain of North Carolina, in the Croatan National Forest, have been established to examine natural regeneration issues. These sites, on moderate to poorly drained soils, contain second-growth longleaf with intact understories, and have been winter burned every two to four years for the last 20 years. We propose to relate stand management, cone production and regeneration rates with measures of site productivity.

INTRODUCTION The Croatan National Forest (CNF) is located on the Coastal Plain of North Carolina between Morehead City and New Bern. Its 65,000 hectares contain longleaf pine communities, bottomland hardwoods, saltwater estuaries, and raised swamps, called pocosins--which account for almost half of the Forest’s acreage. Bounded on three sides by tidal rivers and the Bogue Sound, the forest is defined by water boundaries. Blackwater creeks originate deep in the forest.

Only an estimated 5,000 hectares of longleaf pine (Outcalt and Sheffield, 1996) remain on the CNF, making restoration and management of these communities a priority. The major tools in this effort are the proper use of fire, and utilization of natural regeneration.

Fire plays a crucial role in the management of longleaf communities on the CNF. Fire suppression promotes development of a dense canopy of species other than longleaf pine, growth of a thick shrub midstory, and loss of the fire adapted, shade intolerant understory species. Controlled burns maintain the endangered red- cockaded woodpecker's habitat, while removing competing vegetation for the Croatan's most unusual plants-- the carnivorous Venus fly traps, sundews, and pitcher-plants.

Longleaf pine forests are considered endangered communities of the southeastern USA. Aside from being the dominant tree species, longleaf pine is also considered a keystone species of these community types. Re- establishment and regenerating longleaf pine seedlings is a major focus for conserving and restoring longleaf pine forests throughout the region. Proper management of longleaf pine forests is needed by the National Forest Systems to meet the multiple-use demands placed on that ecosystem, while preserving a rare community type. Another land management consideration is the move toward longer rotations and uneven- aged stands.

STUDY SITES Study sites are located on the Croatan National Forest, Carteret County, NC. All sites are predominantly on Onslow soils--a moderately to somewhat poorly drained, loamy sand (fine-loamy, siliceous, thermic Spodic Paleudults). This soil is highly acidic and generally nutrient poor (Goodwin 1977). Annual precipitation in the region averages 1210 mm but extended droughts occur during the growing season. Mean annual temperature is 17° C with the coldest temperatures in January (0.5° C) and the warmest in July (32.9° C).

The study sites have a typical longleaf flatwoods/savanna vegetative structure--a mature, uneven-aged overstory dominated by longleaf pine (with scattered loblolly and pond pines), no midstory, and relatively low understory dominated by a mix of woody and herbaceous vegetation. Stand ages range from 70 to 100 years. For the past 20 years, study sites have been prescribe burned every 2-4 years in the winter, and all sites have been burned within the last year. In general, the study sites contain similar dominant understory

107 plants, including Gaylussacia spp.; Vaccinium spp.; Ilex spp.; Persea borbonia; Magnolia virginiana; Aristida stricta; Andropogon spp.; Pteridium aquilinum; and Eupatorium spp.

RESEARCH OBJECTIVES AND METHODS To better understand the environmental and vegetative conditions, and management activities that impact cone production, seedling survival, and growth of longleaf pine on poorly drained sites, the following guidelines were established:

• Four study sites, each containing three measurement plots, have been established. • Longleaf pine overstory and grass stage seedlings will be stem mapped. Diameter at breast height (1.37 m; DBH), crown diameter and total height will be measured on the overstory. • Vegetation surveys will be conducted to quantify percent cover and biomass by species. • Longleaf pine flower and cone counts will be done yearly, beginning in 2000. • Estimates of monthly needle production and fuel loads will be determined using litter traps. • Soil temperature, soil moisture, and light (PAR) will be measured monthly. • Soil cores will be taken to determine belowground biomass, and sub samples of soil will be collected for nutrient analyses.

LONGLEAF PINE CONE CROPS One of the major components in longleaf pine management is the production of cones and viable seed. Compared to other southern pines, longleaf is a sporadic seed producer. Wahlenberg (1946) noted that good seed crops might occur every 5 to 7 years, while Maki (1952) reported heavy seed crops occur over much of the longleaf range once in 8 to 10 years.

For successful regeneration, the minimum size of a cone crop is considered to be 1850 cones/hectare or roughly 30 cones per tree (Boyer and White 1989). In the past 30 years, six of the nine cone crops considered adequate for natural regeneration have occurred since 1990 (Boyer 1998). The 1996-longleaf seed crop was one of those “much-anticipated” region-wide seed crops. Whether the interest is natural or artificial regeneration, it is important to know when to expect a bountiful seed crop.

DEVELOPMENT OF LONGLEAF PINE SEEDS The visual development of longleaf pine seed extends into three calendar years. Seed ripen and fall between late September and early November. The following is an abbreviated guideline for the longleaf pine seed development process.

Approximate months prior to seedfall and what happens: 27 months - differentiation between male and female flowers occur; usually July 22 months - male flowers appear, usually December 19 months - female flowers appear and pollination occurs, usually February to April 5 months - fertilization occurs, usually May to June of seedfall year

MANAGEMENT IMPLICATIONS • Monitor cone crops and utilize them when an opportunity occurs. • Larger cone crops are characterized by higher seed viability (Boyer 1973) and increased seedling establishment (Gemmer et al. 1940). • When large crops occur, capture as much reproduction as possible. • Longleaf pine forests can survive reproduction droughts.

CONCLUSION This study will address conditions of regeneration in natural longleaf pine stands. It intends to fill a void of regeneration information on moderately to poorly drained soils of the North Carolina Coastal Plain. This study will provide critical information for the Croatan National Forest in its management and restoration of longleaf pine communities, as well as adding to the body of knowledge about longleaf pine. Ideally, this work will be the first phase of a larger study. Based on the data gathered, we will install treatments to test practices that will provide more predictable and successful natural regeneration within an economically viable framework.

108 LITERATURE CITED Boyer, W.D. 1973. Air temperature, heat sums, and pollen shedding phenology of longleaf pine. Ecology. 54(2):420-426. Boyer, W. D. 1998. Long-term changes in flowering and one production by longleaf pine. In: Waldrop, T.A. ed. Proceedings of the Ninth Biennial Southern Silvicultural Research Conference, USDA Forest Service, Southern Research Station, GTR20: 92-98. Boyer, W.D., and J.B. White. 1989. Natural regeneration of longleaf pine. In Farrar, R.M. ed. Proceedings of the Symposium on the Management of Longleaf Pine. 1989, April 4-6. Long Beach, MS. Gemmer, E.W., T.E. Maki, and R.A. Chapman. 1940. Ecological aspects of longleaf pine regeneration in south Mississippi. Ecological Monographs. 21:75-86.Goodwin, R.A. 1977. Soil survey of Carteret County, North Carolina. Soil Conservation Service. Maki, T.E. 1952. Local longleaf seed years. Journal of Forestry. 50(4):321-322. Outcalt, K.W., and R.M. Sheffield. 1996. The longleaf pine forest: Trends and current conditions. USDA Forest Service, Resource Bulletin SRS-9. 23 p. Wahlenberg, W.G. 1946. Longleaf pine: Its use, ecology, regeneration, protection, growth, and management. Charles Lathrop Pack Forestry Foundation in cooperation with the USDA Forest Service, 429 p.

109 GULF COASTAL PLAIN ECOSYSTEM PARTNERSHIP

Vernon Compton (Gulf Coastal Plain Ecosystem Partnership, 4025 Highway 178, Jay, FL 32565)

ABSTRACT: The Gulf Coastal Plain Ecosystem Partnership (GCPEP) consists of six public and private partners that manage land in the south-central portion of the East Gulf Coastal Plain. The East Gulf Coastal Plain ecoregion covers 42,439,000 acres, stretching from northeastern Louisiana across the southern portions of Mississippi, Alabama, Georgia and western Florida. The exceptional biological diversity in this ecoregion ranks it among the richest in North America. Unfortunately, historical and current habitat loss and alteration also make its biological resources among the most threatened. GCPEP began in part, as a response to the dramatic loss of this longleaf pine habitat across the southeastern United States. The partners manage more than 850,000 acres and include Eglin Air Force Base, the Florida Division of Forestry, the Northwest Florida Water Management District, and National Forests in Alabama, International Paper, and The Nature Conservancy. Contained within the GCPEP landscape are more than 20 percent of the remaining longleaf pine ecosystems and more than 50 percent of the remaining old-growth stands of longleaf pine. These connected lands, however, include more than longleaf pine forests. They contain ecosystems that cut across geographic boundaries, portions of four major watersheds, commercial pine plantations, extensive bottomland hardwood forests, more than 160 rare or imperiled plants and animals and important game and fish populations. Under a Memorandum of Understanding signed in 1996, the partners have agreed to develop and implement a cooperative and voluntary stewardship strategy that sustains native plants and animals, conserves and restores ecosystem integrity, ensures a continued supply of forest commodities, recreational opportunities, clean water and ecosystem services, and supports human communities that depend on these resources and services. To date the partners, along with GCPEP staff, have completed numerous cooperative projects centered on prescribed burning, forest management, ecological restoration, endangered species management and monitoring, and public education. By working together, GCPEP is becoming a leading example showcasing cooperative projects, resulting in even better land and water management, increased public education, and expanded longleaf pine ecosystem recovery.

110 ELEVATED ATMOSPHERIC [CO2] BOOSTS THE PRODUCTIVITY OF MODEL LONGLEAF PINE ECOSYSTEMS WITHOUT ALTERING COMMUNITY STRUCTURE

Micheal A. Davis (School of Forestry and Wildlife Science, 108 M. White Smith Hall, Auburn University, AL 36849; now at Department of Biological Sciences, University of Southern Mississippi, Hattiesburg, MS 39406-5018) Seth G. Pritchard (USDA-ARS Wind Erosion and Water Conservation Research Laboratory, Big Spring, TX 79721) Robert J. Mitchell (Joseph W. Jones Ecological Research Center, Newton, GA 31770) Stephen A. Prior (USDA-ARS National Soil Dynamics Laboratory, 411 S. Donahue Drive, Auburn, AL 36832) Dean Gjerstad (School of Forestry and Wildlife Science, 108 M. White Smith Hall, Auburn University, AL 36849) Hugo H. Rogers (USDA-ARS National Soil Dynamics Laboratory, 411 S. Donahue Drive, Auburn, AL 36832) G.B. Runion (USDA-ARS National Soil Dynamics Laboratory, 411 S. Donahue Drive, Auburn, AL 36832)

ABSTRACT: The current rise in atmospheric [CO2] is projected to double pre-industrial levels within the next century. Doubling atmospheric [CO2] has been shown to increase biomass of C3 plants by an average of 40%. However, individual canopy and understory forest species differ in morphology, physiology, life form and reproductive strategies, and these differences often yield contrasting responses to elevated [CO2]. Differences in species-level responses make predictions of community-level responses to rising CO2 difficult. Also, CO2-induced shifts in competitive advantages between species may alter community composition, structure, and function. To assess the performance of longleaf pine forests to future [CO2], we constructed a model regenerating longleaf pine community composed of species representing different structural and functional guilds: (1) a C3 evergreen conifer (Pinus palustris Mill.); (2) a C4 bunch grass (Aristida stricta Michx.); (3) a C3 broadleaf tree (Quercus margaretta); (4) a C3 perennial herbaceous legume (Crotalaria rotundifolia Walt. ex Gemel); (5) a C3 herbaceous perennial (Asclepias tuberosa L.). Plants were exposed to -1 -1 either ambient (365 μl l ) or elevated (720 μl l ) levels of CO2 within open top chambers. After two years, CO2-enriched plots had 109% greater biomass than ambient plots, mainly due to a 117% increase in pine biomass. An extra growth flush was realized by CO2-fertilized pines in 1999 and 2000. Although pines comprised 4% more of the total biomass in CO2-enriched plots, overall community structure did not change. Our data suggest that longleaf pine will perform well without altering ecosystem structure in a future higher CO2 world.

111 CRP & LONGLEAF PINE: A REGIONAL SUCCESS STORY FOR EXTENSION OUTREACH

Robert M. Franklin (Clemson Extension Service, P.O. Drawer 1086 Walterboro, SC 29488) Dean Gjerstad (School of Forestry, Auburn University, AL 36849) Mark Hainds (Solon Dixon Forestry Education Center, Route 7 Box 131, Andalusia, AL 36420) Rhett Johnson (Solon Dixon Forestry Education Center, Route 7 Box 131, Andalusia, AL 36420)

ABSTRACT: With the 18th Conservation Reserve Program (CRP) sign-up, landowners in the South have an opportunity to restore longleaf pine back to being a major component of southern ecosystems. In Georgia and neighboring states, record numbers of landowners signed up to plant longleaf, but few know the challenge of planting this difficult-to-regenerate species. A total of 102,000 acres were accepted into the CRP program with nearly 75% in Georgia, followed by Alabama (13,480 acres), South Carolina (6,450 acres), Florida (5,968 acres), North Carolina (1,407 acres) and less than 500 acres total in Louisiana, Mississippi, Texas and Virginia.

To assist landowners in learning how to plant longleaf pine, The Longleaf Alliance collaborated with the Alabama Forestry Commission, the Clemson University Cooperative Extension Service, the Georgia Forestry Commission, the Jones Ecological Research Station, USDA Forest Service and the USDA Natural Resources Conservation Service to do a series of longleaf pine tree planting workshops in the southeast. Between January and October of 1999, thirteen workshops were held in Alabama, Georgia and South Carolina, the states with the largest CRP longleaf enrollment.

Over 1,000 landowners and natural resource professionals received training in: Longleaf Pine Ecosystems, Use of Prescribed Burning, Site Preparation Options, Care & Planting of Seedlings, Economics of Longleaf Pine and participated in a planting demonstration that featured hand and machine planting with the proper equipment.

Based on written evaluations of six of the workshops, 202 landowners rated the programs highly and indicated they would plant over 41,200 acres of longleaf pine over the next two years. They indicated that the knowledge gained in planting and management would help them earn $1.75 million dollars and save $930,000 dollars in the future when working with longleaf pine.

Plans are underway to update the workshop and offer it again this year for the next group of landowners enrolling in the CRP longleaf program. The program materials are available from The Longleaf Alliance to anyone within the region with an interest in planting longleaf pine. For additional details on this and other longleaf-related outreach efforts, contact The Longleaf Alliance at (334) 222-7779.

112 THE SOUTHEASTERN ECOLOGICAL FRAMEWORK: BUILDING ECOLOGICAL AND ORGANIZATIONAL CONNECTIVITY THROUGHOUT THE SOUTHEAST

Stephanie Fulton (US EPA/OPM-Planning, 61 Forsyth Street, Atlanta, GA 30303)

ABSTRACT: EPA is always looking for ways to partner with communities to achieve programmatic goals. EPA Region 4 has developed a tool to aid communities and organizations in targeting and leveraging funds for natural resource protection. Most conservation efforts have been geared at protecting individual species and small, local ecosystems. Piece-meal protection of the environment, however, often leads to degradation of the parts being protected. The resulting fragmentation prevents the operation of many large-scale processes from adequately functioning. The more we chop up the natural environment into smaller and smaller isolated pieces, the less able are these systems to continue to function as whole systems. While it is important to protect critical habitats and watersheds, it is equally as critical to protect ecological corridors that connect these habitats through remaining semi-natural areas. We have developed a regional prototype, the Southeastern Ecological Framework, which is essentially a regional? greenprint? that can be used by anyone interested in protecting water quality, species habitat, important ecological areas, quality of life and other important natural features by preserving connectivity between those natural areas. The framework is comprised of important regional ecological hubs and "greenways" corridors that connect them through out the 8-state region. This greenways corridor could be used as a planning or mitigation tool to target areas for longleaf pine restoration that would not only help restore the species in its native habitat, but help preserve or restore ecological connectivity in the landscape.

113 LONGLEAF PINE REGENERATION DYNAMICS IN ARTIFICIALLY CREATED GAPS IN THE APALACHICOLA NATIONAL FOREST, FLORIDA

J.L. Gagnon (School of Forest Resources and Conservation, University of Florida, 370 Newins-Ziegler Hall, P.O. Box 110410, Gainesville, FL 32611-0410) E.J. Jokela (School of Forest Resources and Conservation, University of Florida, 370 Newins-Ziegler Hall, P.O. Box 110410, Gainesville, FL 32611-0410) W.K. Moser (Tall Timbers Research Station, 13093 Henry Beadel Road, Tallahassee, Florida 32312-0918) D. Huber (School of Forest Resources and Conservation, University of Florida, 363 Newins-Ziegler Hall, P.O. Box 110410, Gainesville, FL 32611-0410)

ABSTRACT: The group selection system has gained popularity as a means for developing and maintaining uneven-aged structure in longleaf pine stands. Few studies documenting longleaf pine seedling development within canopy gaps are available as a basis for management decisions. The objectives of this study were to test whether seedling survival and growth varied in relation to canopy gap position and site resource availability. Four, 0.2-ha circular gaps were cut in a previously unmanaged 70-yr-old natural longleaf pine stand. Within each gap, four plots were established at each of three positions representing different levels of edge effect competition (i.e., center of the gap (C1), middle (C2), and edge (C3)). Each plot was subdivided and two treatments were randomly assigned (control vs. fertilizer + hand weeded) to examine the interactions between canopy gap position and site resource availability on regeneration dynamics. Containerized longleaf pine seedlings were planted in 25 tree plots (n=24 plots per gap) in February 1999 and monitored for survival and root collar diameter (RCD). PAR levels and soil moisture were determined throughout the growing season. After one growing season, differences in within-gap survival (23% (C1) to 51% (C3)) and RCD growth (2.6 mm (C1) to 1.6 mm (C3)) were apparent. A negative treatment effect was observed on survival (41% control, 31% treated); the opposite was true for growth (RCD = 1.7 mm control, 2.9 mm treated). Although no differences in soil moisture were found among gap positions, PAR levels were highest in C1 (96% of incident) and lowest in C3 (72% of incident).

114 LONGLEAF PINE THREE YEARS AFTER PLANTING AT NAVAL AIR STATION PENSACOLA, FLORIDA

Mark W. Gibson (Navy Regional Command Pensacola Florida, Code 22000, 190 Radford Blvd., Pensacola, FL 32508-5217)

ABSTRACT: Longleaf pine (Pinus palustris) bareroot seedlings were hand planted on a 30-acre site in January 1998. The planting site was clearcut in the spring of 1997 and site prepared by drum chopping in November 1997. Genetically improved seedlings from the Florida Division of Forestry nursery in Chiefland, Florida, were planted with a "hoe-dad" on a 6' x 10' spacing.

The site occurs on a broad ridge top of an old dune front, one-quarter mile north of the Intracoastal Waterway, and one mile north of the Gulf of Mexico, in Escambia County of the Florida Panhandle. Site elevation is approximately 25' above msl. Soils on the site are excessively well-drained, deep sands, Kureb Series. Site vegetation prior to reforestation consisted of a scattered sand pine (Pinus clausa) overstory with a continuous midstory and understory "scrub" of oak and shrub species and a discontinuous herbaceous forest floor with open sandy areas. In the early 1950's, the site was clearcut prior to Navy ownership, removing all longleaf pine (and seed source) that occurred on the site allowing succession to sand pine. Longleaf and sand pine often compete on sandy coastal sites where their ranges overlap, with sand pine often "winning" the competition, especially if fire is restricted. In 1995, Hurricanes Erin and Opal caused significant damage to the sand pine overstory, leading to the decision to regenerate the site and to re-establish longleaf pine.

Three years after planting, exceptional longleaf pine growth has been observed. Several individual seedlings came out of the grass stage early in the second season and are now 4' - 6' in height. This type of growth has not been observed at other longleaf plantings on Navy lands. Precipitation of 23.3 inches for January and February during the year of planting was three times above normal, and the yearly total was 10 inches above normal. For the two years following planting, a precipitation deficit of nearly 50 inches has occurred. Seedling genetics in response to these extremes could explain some of the exceptional growth. Additionally, the early and exceptional height growth occurred only in open sandy areas of the site where competition was not a factor. In other areas of the site where site preparation methods were not sufficient to control competing vegetation, severe mortality and below normal growth occurred.

115 LONG-TERM STUDIES ON DEVELOPMENT OF LONGLEAF PINE PLANTATIONS

J.C.G. Goelz (USDA Forest Service, Southern Research Station, 2500 Shreveport HWY, Pineville, LA 71360) D.J. Leduc (USDA Forest Service, Southern Research Station, 2500 Shreveport HWY, Pineville, LA 71360)

ABSTRACT: The U.S. Forest Service's Laboratory at Pineville, LA has established and maintained over 250 permanent plots in longleaf pine plantations. This database represents a range of sites in south-central United States. Some of these plots represent over 50 years of stand development in plantations currently over 65 years old. All of the plots have recorded 20 years or more of stand development. These plots are being used to develop a model to predict the growth and development of longleaf plantations. This model will be a valuable tool for managers of longleaf plantations. In addition to these valuable, historic plots, the Pineville laboratory is currently establishing new permanent plots in young longleaf pine stands, representing operational plantations on a variety of sites.

The poster will describe the individual studies comprising the longleaf pine database, and will provide growth trends observed on some individual plots. Furthermore, this will provide an opportunity to initiate cooperation of the Pineville laboratory with others interested in the development of longleaf pine plantations.

THE LEGACY For over 50 years, the USDA Forest Service’s Laboratory at Pineville, LA has been conducting research on the growth and development of longleaf pine plantations. We have seven long-term studies scattered throughout the Gulf Coastal Plain. These studies have explored the effects of spacing and thinning on longleaf plantations; most of the studies have been thinned several times. Each of the studies has recorded 20 years or more of stand development in longleaf pine plantations. Our database consists of 267 permanent plots.

THE LEGACY CONTINUES Most of our plots were established more than 20 years ago. While this provides us with an invaluable long- term database, we have lost the ability to determine how applicable this data is to current cultural and environmental factors, as current conditions are not reflected. Therefore, we are initiating a new study that will establish permanent plots in young, operational plantations (this work will involve not only longleaf, but also loblolly and slash pines). Given sufficient support and cooperation, this work will embrace a wide range of sites, practices, and geography. We seek cooperators reflecting all ownerships of forestlands including governmental agencies, industrial forest landowners, and non-industrial forest landowners. This data will allow development of forest models that reflect the range of current practices.

FULLFILLMENT OF THE LEGACY Although the data represent an interesting record of the development of longleaf pine plantations, the ultimate benefit from this legacy of effort to establish, maintain, and regularly remeasure these plots will be obtained in the construction of a model that will describe the growth and development of longleaf plantations for specified site quality, spacing and thinning. A growth and yield model for unthinned plantations was developed about 20 years ago (Lohrey and Bailey 1977). Since that time, most of our plots have been remeasured several times. This more-recent data will provide much more evidence concerning how older plantations develop. Bernie Parresol of the Biometrics unit at Asheville has conducted some further modeling work emphasizing unthinned plantations and height growth models. Dan Leduc of our laboratory has conducted some interesting work in using neural networks to model diameter distributions of unthinned longleaf plantations (Leduc et al. 1999). The “grass stage” that is often exhibited by longleaf pine provides an interesting challenge to modeling diameter distributions; often the diameter distribution is bi- or multi- modal. This bimodality may be maintained well past the grass stage, as longleaf saplings and poles often persist in subordinate crown positions. We plan to complete an entire growth and yield model for longleaf pine around this time next year. This will allow managers to project the growth and development of longleaf plantations, including the effects of initial spacing, thinning, and site quality. As our database is limited to stands of less than 70 years, and this is less than the planned rotation age of longleaf for some managers, we plan to constrain our model so that it will give reasonable results for projections up to age 100 or 120. We anticipate subsequent modeling efforts that will incorporate the effects of establishment techniques/stock

116 types, fire, and fertilizer, and predict understory community development, and also provide a merchandising processor that will segregate volume into several product classes, including poles.

FORETASTE OF THE LEGACY In anticipation of the results of our more-formal modeling efforts, we present yield tables for unthinned plantations. These yield tables (tables 1-3) represent the average per-acre values across all our plots. Thus, the progression of values over age or across site are not necessarily smooth. The tables were produced by calculating mean values for each age and site class. Overall, we had 877 observations. The cell for medium sites at age 20 is based on 66 observations; this is the most for any cell. The cell based on the fewest observations is for medium sites at age 5; only 5 observations were available. We feel the table provides a “reality check” for expectations of yield of longleaf pine plantations. Although these plantations were established before current establishment practices for longleaf were devised, the plantations represent well- cared-for forests whose establishment was much more prompt than many plantations that were established at the time. Thus, we suspect that this table may adequately reflect the expectations for current operational plantations.

Table 1. Average Number of Trees Per Acre Site Class Low Medium High Age Class SI25 < 50 50 >SI25 < 60 SI25> 60 5 569 786 10 427 722 15 326 510 815 20 248 466 718 25 158 565 591 30 136 530 514 35 124 477 442 40 128 363 384 45 120 320 293 50 89 279 215 55 87 242 60 86 187 65 82 169 Table 2. Basal Area (ft2) per acre Site Class Low Medium High Age Class SI25 < 50 50 >SI25 < 60 SI25> 60 5 5 14 10 23 40 15 26 59 109 20 39 82 129 25 44 118 150 30 55 140 166 35 62 156 174 40 77 157 178 45 85 164 163 50 88 163 142 55 94 172 60 100 171 65 104 177

117 Table 3. Volume (total ft3 outside bark) per acre Site Class Age Class Low Medium High SI25 < 50 50 >SI25 < 60 SI25> 60 5 43 156 10 316 581 15 412 1104 2304 20 770 1853 3280 25 956 3057 4331 30 1348 4175 5372 35 1711 5049 6097 40 2328 5375 6556 45 2724 5941 6243 50 2993 6166 5640 55 3328 6710 60 3647 6808 65 3887 7234

LITERATURE CITED Leduc, D.J., T.G. Matney, and V.C. Baldwin, jr. 1999. Diameter distributions of longleaf pine plantations – a neural network approach. In: Kush, J.S., comp. 1999. Longleaf Pine: A Forward Look. Proceedings of the 2nd Longleaf Alliance conference; 1998 November 17-19; Charleston, SC. Longleaf Alliance Report No. 4. p 112-114. Lohrey, R.E., and R.L Bailey. 1977. Yield tables and stand structure for unthinned longleaf pine plantations in Louisiana and Texas. USDA Forest Service, Southern For. Expt. Stn. Res. Pap. SO-133. 53 p.

118 SPATIO-TEMPORAL DYNAMICS OF POND USE AND RECRUITMENT IN FLORIDA GOPHER FROGS (RANA CAPITO AESOPUS)

Cathryn H. Greenberg (USDA Forest Service, Southern Research Station, Bent Creek Experimental Forest, 1577 Brevard Road, Asheville, North Carolina 28806)

ABSTRACT: This study examines spatio-temporal dynamics of Florida gopher frog (Rana capito aesopus) breeding and juvenile recruitment. Ponds were situated within a hardwood-invaded or a savanna-like longleaf pine-wiregrass upland matrix. Movement (n = 1,444) was monitored using intermittent drift fences with pitfall and funnel traps at eight isolated, ephemeral ponds February 1994-January 1999. Adult pond use was low but similar among years and habitat matrices. Juvenile recruitment was higher in the savanna-like upland matrix. The number of adults using ponds was positively correlated with the number of next-years recruit in only one year. Recruitment rates were relatively low (maximum 175 captured/pond/yr), but juveniles were produced from most ponds in three of five years. Recruitment was negligible in 1994 due to ponds drying and in 1997 for unknown reasons. Juvenile body size differed among years and ponds but was negatively correlated with the number of juveniles exiting ponds in only one year, suggesting that intraspecific competition is only one of many factors affecting juvenile body size. Most emigration by juveniles occurred May-August and was unrelated to rainfall. High variability in juvenile recruitment success and significant differences in body size among years and ponds suggests that each is influenced by factors at both a landscape (e.g., rainfall and pond hydrology) and within-pond scale (e.g., competition and predation).

119 COMPARISON OF SITE PREPARATION METHODS AND HERBACEOUS RELEASES FOR LONGLEAF PINE (PINUS PALUSTRIS) ESTABLISHMENT IN AN OLD PECAN ORCHARD

M. J. Hainds (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849) D. H. Gjerstad (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849) E. E. Johnson (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849)

ABSTRACT: Approximately 100,000 acres of agricultural lands across the SE- U.S will be planted to longleaf pine (Pinus palustris) through the Conservation Reserve Program (CRP) during the 1999-2000 planting season. Many additional acres will be planted over the next few years through the CRP. Successful establishment of southern pines on agricultural lands has proven difficult, and successful longleaf pine establishment can be especially challenging. The most difficult sites to successfully plant appear to be agricultural lands with significant components of grass species. This study compares 3 site preparation methods: subsoiling only, scalping & subsoiling, and chemical site preparation with a tank-mix of Roundup Ultra (glyphosate) and Arsenal (imazpyr) plus subsoiling on an old pecan orchard with many grass and broadleaf weed species present including; Cynodon dactylon, Paspalum notatum, and Digitaria spp. This study also examines the effects of 11 different herbaceous releases on longleaf pine following the afore- mentioned site preparations. Scalping was found to be particularly effective in successfully establishing longleaf pine seedlings. Some herbaceous releases appeared very promising, especially those containing Oust (Sulfometuron) either as a stand-alone treatment, or as a tank-mix with Velpar (Hexazinone) or Arsenal (Imazapyr). Correct timing of the herbaceous release is an important factor when minimizing damage to longleaf pine seedlings and controlling herbaceous competition.

120 CONTAINER-GROWN LONGLEAF (PINUS PALUSTRIS) SEEDLING SURVIVAL AT FOUR PLANTING DEPTHS

M. J. Hainds (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849) E. E. Johnson (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849) D. H. Gjerstad (Longleaf Alliance and the School of Forestry and Wildlife Sciences, Auburn University, AL 36849)

ABSTRACT: Container-grown longleaf pine seedling production has increased from approximately 30 million seedlings in 1996 to approximately 80 million in 2000. Previous planting experiences led the authors to believe that precise planting depth is critical for the successful establishment of both bareroot and container-grown longleaf pine seedlings. However, a literature review revealed no studies examining survivorship rates for seedlings planted at varied planting depths. To evaluate the effect of planting depth on seedling survival, a study was initiated using container-grown longleaf seedlings planted the second-week of December 1998. Seedlings were hand-planted with a “plug tool” in sandy loam soils on a previously scalped and subsoiled (ripped) site. Seedlings were planted at four depths:

• Exposed plug, 1 cm above soil surface. • Exposed plug, planted at soil surface. • Plug planted 1 cm below soil surface, plug covered (terminal bud not covered). • Plug planted 2 cm below soil surface, plug covered (terminal bud covered).

Seedling survival was initially evaluated on 5/18/99. Seedling survival and growth were re-measured on 7/19/00

121 COMPARISON OF RED-COCKADED WOODPECKER NESTLING DIET IN OLD-GROWTH AND OLD-FIELD LONGLEAF PINE HABITATS

James L. Hanula, (USDA Forest Service, Forestry Sciences Laboratory, 320 Green Street, Athens, GA 30602) R. Todd Engstorm, (Tall Timbers Research Station, 13093 Henry Beadel Road, Tallahassee, Florida 32312)

ABSTRACT: Automatic cameras were used to record adult red-cockaded woodpecker (Picoides borealis) nest visits with food for nestlings. Diet of nestlings on or near an old-growth longleaf pine (Pinus palustris) remnant in southern Georgia were compared to that in longleaf pine stands established on old farm fields in western South Carolina. Diets of nestlings were expressed as percent nest visits and percent prey biomass. The method of calculating nestling diet composition had little effect on the relative ranking of prey. Roaches (Blattaria: Blatellidae) were the most common fed to nestlings, ranging from 33-57% of the prey brought to nest cavities by adults or 55-73% of the prey biomass. Other common prey were spiders, centipedes and caterpillars. Morisita's index (C) of diet overlap showed a high degree of similarity in nestling diets among years in the old-growth remnant (C = 0.91 to 0.94), as well as a high degree of similarity in the diets of nestlings among woodpecker groups within locations and between old-growth and old-field habitats (C = 0.89-0.95). Our study shows that old trees on relatively undisturbed sites with natural understory plant communities provide the same prey as younger trees growing on old farm fields and the relative importance of the different prey was similar for both habitats.

For more information see: Hanula, J. L. and R. T. Engstrom. in press. Comparison of red-cockaded woodpecker (Picoides borealis) nestling diet in old-growth and old-field longleaf pine (Pinus palustris) habitat. Am. Midl. Nat. 144.

122 EFFECTS OF 40 YEARS OF WINTER BURNING IN LONGLEAF PINE ON INSECTS AND OTHER ARTHROPODS

James L. Hanula (USDA Forest Service, Southern Research Station, 320 Green St., Athens, GA 30602) Dale D. Wade (USDA Forest Service, Southern Research Station, 320 Green St., Athens, GA 30602)

ABSTRACT: Longleaf pine (Pinus palustris) once occupied 60-90 million acres but it is found on <3 million acres today. Much of the remaining longleaf habitat no longer contains the representative understory plant communities, so recent efforts are focusing on using fire to restore them. However, little is known about the long-term impacts of this strategy. A unique study on the Osceola national Forest in northern Florida allowed us to determine the long-term effects of winter prescribed burning applied at frequencies of 1, 2 or 4 years over a 40 year period on the abundance and diversity of insects and other arthropods. Plots burned annually had the lowest diversity while plots burned every 2 or 4 years were similar. All three treatments reduced arthropod diversity below that of unburned controls. Burning did not increase the number of rate species but arthropod community composition were affected. Over 100 species were affected by burning. Spiders as a group were the most severely affected but arthropods in 11 other orders were also impacted by winter burning. Some populations responded positively to burning while others were reduced. In many cases, 4 years was not enough time for arthropod populations reduced by burning to recover to the levels found on unburned controls. The results suggest that some areas should be left unburned to remain overall diversity on the landscape.

Longleaf pine, Pinus palustris, forests once occupied >24 million hectares in the southern United States. Today, <1.3 million hectares remain as small isolated parcels (Outcalt and Sheffield 1996). Of those, less than 1 million retain the longleaf pine/grassland association. Although most agree that increasing longleaf pine abundance in the landscape and restoration of existing degraded longleaf pine communities is a desirable goal, how to achieve that goal is unclear. Under natural conditions longleaf pine communities probably burned during the growing season and growing season fires are recommended for longleaf pine community maintenance and restoration (Frost 1990). However, burning large acreages during the growing season at recommended fire frequencies is difficult because the growing season provides fewer days when wind and fuel conditions are good and new Environmental Protection Agency restrictions on smoke and volatile emissions from prescribed burning may prevent significant increases in growing season burning. Therefore, dormant season prescribed burning is likely to remain an important management tool for longleaf pine communities. We took advantage of a unique long-term study on the Osceola National Forest (Baker Co., FL) where growing season burns were applied to longleaf pine stands over a 40 year period at frequencies of 1, 2, or 4 years. The study was replicated (N=6) and included unburned control plots. The study was conducted from November 1994 to October 1999.

We measured arthropod abundance with pitfall traps and termite abundance with wooden trap blocks. We also measured large woody debris volume, log decomposition and nitrogen content, and live and dead plant material on the plots. Logs were placed on the plots in November 1994 and sampled annually (1 log/year/plot) to measure decomposition rates (change in specific gravity) and changes in nitrogen content of the wood. Wooden blocks (5x10x25cm) were distributed over the plots (15/plot) and monitored every other month for termites over a two-year period.

We operated 8 pitfall traps/plot every other month throughout the study. Four pitfalls were located along 3 m long logs and 4 were along 3 m long metal drift fences. We did not see any interaction between arthropod use of logs and burn frequency so we combined the two types of traps for analyses of burn effects on arthropods.

The frequency of burning had no effect on the overall amount of large woody debris on the plots or the numbers of snags but it did affect the volume of logs. Plots burned annually or biennially had lower log volumes than plots burned quadrennially. The rate of log decomposition and the loss of nitrogen from wood was unaffected by burning frequency. Winter burning frequency also had no effect on termite abundance. These results suggest that winter burning has little effect on the wood decomposer community of longleaf forests but that winter burning does reduce large woody debris lying on the forest floor.

123 Burning altered the composition of the live and dead plant material on the forest floor. Total dead plant matter was almost 5 times higher on the unburned control plots than on the annually burned plots. Likewise, live plant biomass was over 4 times higher on the unburned plots. Annual winter burning caused nearly a 20- fold reduction in palmetto biomass and a 5 fold increase in grasses. Gallberry was unaffected.

We looked at over 163,000 arthropods from, 31 orders, 265 families, and 932 genera. We observed an overall increase in total arthropod abundance and biomass in burned plots compared to unburned ones but this was do to one very abundant millipede. Frequent winter burning reduced the diversity of ground-dwelling arthropods but did not affect overall richness or the number of rare genera (<5 caught/5years). Frequent burning reduced arthropod community similarity.

We classified each arthropod into one of four groups by feeding habits. Dormant season burning reduced predators compared to unburned controls. Detritus feeders increased in abundance with burning, herbivore numbers were the same on all treatments, and omnivores as a group were also relatively unaffected by dormant season burning.

A total of 86 different arthropods were affected by dormant season burning in some way. Frequent burning reduced the abundance of 41 genera and in many cases 4 years was not enough time for their populations to recover. Thirty-one genera increased with frequent burning and 14 genera exhibited a response to intermediate frequencies of burning.

These results are the first to show how burning over a long period of time affects arthropods. The slow recovery rate of so many species suggests that management oriented toward conservation of biodiversity in longleaf pine should include areas excluded from fire. Our plots were less than 1 ha but they apparently were large enough to provide a refuge for a number of species. In addition, the unburned plots were invaluable for research allowing us to detect effects on a number of species that otherwise would have gone unnoticed without them.

REFERENCES Beach, V. 1993. An up & coming forest. South Carolina Wildlife. Jan/Feb: 44-49. Frost, C. C. 1990. Natural diversity and status of longleaf pine communities. In: Forestry in the 1990's - a changing environment. Proc. Society of American Foresters Reg. Tech. Conf. Pinehurst, NC. Outcalt, K. W., and R. M. Sheffield. 1996. The longleaf pine forest: trends and current conditions. U.S. Forest Service Resource Bull. SRS-RB-9.

124 VERTEBRATE COMMUNITY RESPONSE TO ALTERNATIVE APPLICATIONS OF PRESCRIBED FIRE IN LONGLEAF PINE FORESTS OF THE NORTH CAROLINA SANDHILLS: PRELIMINARY RESULTS

Carol L. Hardy (USDA Forest Service, National Forests in Mississippi, 100 W. Capitol St., Jackson, MS 39269) Leonard A. Brennan (Tall Timbers Research Station, Inc., 13093 Henry Beadel Dr., Tallahassee, FL 32312) Sudie E. Daves (Tall Timbers Research Station, Inc., 13093 Henry Beadel Dr., Tallahassee, FL 32312-9712) L. Wesley Burger (Department of Wildlife and Fisheries, P.O. Box 9690, Mississippi State University, Mississippi State, MS 39762)

ABSTRACT: Prior to European settlement, longleaf pine and other southern pine ecosystems dominated the southeastern coastal plain. Today, naturally-regenerated longleaf pine is one of the most imperiled natural communities in the southeastern United States and is therefore of special conservation concern. Prescribed fire during the lightning season may better mimic natural fires that historically shaped the development and maintenance of this system. This study tests the effects of seasonal applications of prescribed fire on the distribution and abundance of vertebrates in longleaf pine forests. The study was conducted on Fort Bragg Military Installation and the Sandhills Game Lands in the Sandhills region of North Carolina. Four sets of paired sites were located in mature upland longleaf sites. Treatments were biennial applications of dormant- season and lightning-season prescribed fire. Experimental design encompassed pre-treatment surveys and post-treatment assessments of avian, small mammal, and terrestrial herpetile species. Portions of the invertebrate and plant community were also characterized. We present results of the first treatment cycle (1996-1998). Avian community species richness and number of territories did not differ between dormant- and lightning-season sites. Avian nest success declined on both treatments during the year of burn, then increased during the 2nd growing season post-burn. Small mammal numbers declined on both treatment sites during year of burn and did not increase the following year. Herpetile species richness and individual abundance were more influenced by rainfall amount than by season of fire. Mean weights for some herpetile species declined the year of burn on both treatment sites and increased during the 2nd growing season post- burn. Insect ordinal richness and number of individuals did not differ between treatments, but average biomass declined on both treatment sites the year of burn and increased the following year. Percent understory hardwood stems top-killed were higher on lightning-season than dormant-season sites. Percent ground cover characteristics (wiregrass, leaf litter, etc) did not differ between treatment sites during this first fire cycle. A delayed recovery rate of understory vegetation following both fire treatments was observed. This could be a function of low soil fertility as compared to other regions. Delayed vegetation response could account for observed decreases in insect biomass, avian reproductive success, small mammal abundance, and herpetile weight in both treatments. In general, differences between treatments in plant and vertebrate communities were much more subtle than was temporal variation in relation to application of either fire regime. Insofar as different long-term fire regimes produce different plant communities, a divergence in vertebrate communities, not detectable in this first fire cycle, might be expected over several fire cycles.

INTRODUCTION Prior to European settlement, longleaf pine (Pinus palustris) and other southern pine ecosystems dominated the southeastern coastal plain. Today, naturally-regenerated longleaf pine is one of the most imperiled natural communities in the southeastern United States and is therefore of special conservation concern. Prescribed fire during the lightning season may better mimic natural fires that historically shaped the development and maintenance of this system. This study tests the effects of seasonal applications of prescribed fire on the distribution and abundance of vertebrates in longleaf pine forests. The study was a cooperative effort between the Department of Defense, North Carolina Wildlife Resources Commission, U.S.D.A. Forest Service, Tall Timbers Research Station, Inc., and Mississippi State University, and was conducted on Fort Bragg Military Installation and the Sandhills Game Lands in the Sandhills region of North Carolina. We present results of the first treatment cycle (1996-1998).

METHODS Four sets of paired 12-ha study sites were established in mature (>55 years) longleaf pine forests; two sets on Fort Bragg and two sets on the Sandhills Game Land. Each set of paired sites were randomly assigned a dormant- or lightning-season prescribed fire treatment. Fire treatments were applied biennially (every other

125 year) in 1997 and 1999. Pre-treatment surveys and post-treatment assessments of terrestrial communities were completed on both treatment sites, for a total of five years of data collection. Breeding and wintering bird community variables were measured using a combination of spot-mapping, nest searches, behavioral scoring, and repeat census techniques (Kolb 1965, Robbins 1970, Hensler 1985, Vickery et al. 1992, Martin and Geupel 1993). Vegetation structure and composition around each nest located was described according to a modification of nest-searching protocol (Anonymous 1994). Small mammal community variables were assessed using grids of Havahart live traps and mark-recapture techniques (Seber 1982). Arthropod communities associated with groundcover vegetation were sampled using a hand-held D-Vac motorized vacuum collector along 30m transects within study sites (Hurst 1972). Four additional mature longleaf pine sites containing known natural ephemeral ponds were selected for herpetile sampling. The upland area surrounding each pond was divided in half, and each side randomly assigned a prescribed fire treatment (the pools were not burned). Pitfall trapping with drift fences were used to determine amphibian and reptile community variables (Means and Campbell 1981, Campbell and Christman 1982). Overstory, midstory and coarse woody debris variables were measured annually on ten 0.1 ha subplots within the 12ha study sites (Thomas et al. 1979, National Park Service 1992). Groundcover structural characteristics were measured biannually (winter and summer) on both the 12ha and ephemeral pond study sites using a dowel-intercept method.

STATISTICAL ANALYSIS For this initial analysis, avian community variables were analyzed using a combination of ANOVAs using Mixed Models and Z-test procedures (Steel and Torrie 1980, Littell et al. 1996). 12ha study sites were set up as a randomized block design with paired replicates. Ephemeral pond study sites were set up as a randomized block design with ephemeral ponds acting as replicates. In Mixed Models ANOVAs, blocks were treated as random effects and treatments as fixed effects.

RESULTS Avian Community - breeding No differences in species richness were observed between dormant- and lightning-season fire treatments (Range 29-39; p > 0.14). The number of territories of species by nest guild (ground/shrub, cavity, and tree nesting) also did not differ between treatments (Ground/shrub R 5-14.75, p > 0.53; Cavity R 5-4.75, p > 0.20; tree R 8.5-19.25, p > 0.17). However, the number of territories of ground/shrub and tree nesting species increased the year following fire for both treatments. Percent success of cavity-nesting bird species was relatively constant between treatments (65-92%). Percent success of open-cup nesting bird species declined the year of fire (1997) for both treatments, then returned to previous levels (‘96, 30-60%; ‘97, 13-34%; ‘98, 40-52%; p = 0.0001)

Avian community - winter Species richness was higher on dormant-treated sites the year of fire (dormant 19 vs lightning 13.5; p = 0.026). The average number of individuals per site did not differ between treatments. (R 48-228; p >0.24).

Small mammal community A higher average number of individuals were caught on Ft. Bragg sites than on Sandhills Game Land sites (21.5 vs 2.5; p = 0.039). The average number of individuals captured declined the year of fire treatment for both treatments and remained lower the following year (‘96, 21/2; ‘97 9/1; ‘98, 11/1).

Arthropod community No differences in ordinal richness or number of individuals captured were observed between treatments (richness R 9.5-11, p > 0.31; individuals R 338-696, p > 0.29). There was no difference in average total biomass (g) between treatments, but there was a significant decrease in average total biomass for both treatments sites the year of fire treatment, followed by a subsequent increase in average biomass the year following fire (‘96, R 0.006-0.0143; ‘97 R 0.00386-0.004744; ‘98 R 0.0148-0.0536; p = 0.0023).

Herpetile community No differences were observed in amphibian or reptile species richness between treatments (amphibian R 7- 10, p > 0.54; reptile R 4-8, p > 0.04). There were also no differences in amphibian or reptile average number of individuals captured between treatments (amphibian R 47-254, p > 0.57; reptile R 7-23, p > 0.46).

126 Average weights (g) of common species such as southern toads (Bufo terrestris) and eastern spadefoot toads (Scaphiopus holbrookii) did not differ between treatments, but there was a decrease in average weights of both species the year of fire for both treatments, followed by an increase the year following fire treatment (southern ‘96, R 33.35-35.96; ‘97 22.76-23.24; ‘98 25.09-27.39. Eastern spadefoot ‘96 R 37.09-41.0; ‘97 R 33.87-34.68; ‘98 R 35.17-39.25).

Vegetation community A higher percentage of understory hardwood stems were top-killed on lightning-season fire treatment sites (lightning R 54-76% vs dormant R 28-25%; p = 0.02). However, there were no differences in percent ground cover characteristics (% of ground covered by wiregrass, litter, shrub, etc.) between fire treatments (Wiregrass R 21-27%; litter R 38-59%; shrub R 0-7%; forb R 4-8%; legume R 1-5%; bare ground R 1-7%; p > 0.17). There was a slower than expected re-growth of understory ground cover following both fire treatments.

CONCLUSIONS AND MANAGEMENT IMPLICATIONS Vertebrate community variables measured responded similarly to prescribed fire treatments. The observed decrease in insect biomass in both lightning and dormant-season treatment sites the year of fire (1997) may be due to slow vegetation re-growth due to poor, sandy soils. Decreases in nest success of avian open-cup nesting species in both treatments the year of fire may be due to a combination of slow vegetation re-growth and lower insect biomass. The observed decline in average weight of some amphibian species in both treatments the year of fire might also be related to lower insect biomass. Small mammal populations may take more than one year to recover from both lightning- and dormant-season fire treatments to pre-fire levels. In general, differences between treatments in plant and vertebrate communities were much more subtle than was temporal variation in relation to application of either fire regime. Insofar as different long-term fire regimes produce different plant communities, a divergence in vertebrate communities, not detectable in this first fire cycle, might be expected over several fire cycles.

ACKNOWLEGDMENTS The authors would like to thank the Department of Defense, Ft. Bragg Military Installation's Wildlife Branch, Endangered Species Branch and Natural Resources Branch and the North Carolina Wildlife Resources Commission and Sandhills Wildlife Depot for financial, logistical and moral support during this project. We would also like to thank the USDA Forest Service Southern Research Station's Longleaf Pine Restoration Project for financial support of this project.

LITERATURE CITED Anonymous. 1994. BBIRD field protocol: breeding biology research and monitoring database. Montana Cooperative Wildlife Research Unit, University of Montana, Missoula, MT. Campbell, H.W., and S.P. Christman. 1982. Field techniques for herpetofaunal community analyses.193-200 in N.J. Scott (ed.) Herpetological Communities. Research Paper 13. U.S. Fish and Wildlife Service, Washington D.C. Hensler, G.L. 1985. Estimation and comparison of functions of daily nest survival probabilities using the Mayfield method. Pp. 289-301 in B.J.J. Morgan and P.M. North, eds. Statistics in ornithology. Springer- Verlag, Berlin. Hurst, G.A. 1972. Insects and bobwhite brood habitat management. Proceedings of the National Bobwhite Quail Symposium 1:65-81. Kolb, H. 1965. The Audubon winter bird-population study. Aud. Field Notes 19:432-434. Littell, R.C., G.A. Milliken, W.W. Stroup, and R.D. Wolfinger. 1996. SAS system for mixed models. SAS Institute, Inc. Cary, NC. 633 pp. Martin, T.E. and G.R. Geupel. 1993. Nest-monitoring plots: methods for locating nests and monitoring success. J. Field Ornithol. 64:507-519. Means, D.B. and H.W. Campbell. 1981. Effects of prescribed burning on amphibians and reptiles. Pages 89-98 in G.W. Wood, (ed.) Prescribed Fire and Wildlife in Southern Forests. Belle W. Baruch Forest Science Institute, Georgetown, S.C. National Park Service. 1992. Western Region Fire Monitoring Handbook. USDI National Park Service, Western Region, San Francisco, CA. Robbins, L.E., and R.L. Myers. 1992. Seasonal Effects of Prescribed Burning in Florida: a Review. Tall

127 Timbers Research, Inc; Miscellaneous Publication No. 8., Tallahassee, FL. Seber, G.A. 1982. The estimation of animal abundance and related parameters. Griffen, London, England. Steele, R.G. and J.H. Torrie. 1980. Principles and procedures of statistics: a biometric approach. McGraw-Hill, Inc. 623 pp. Thomas, J.W., R.G. Anderson, C. Maser, and E.L. Bull. 1979. Snags. Pg 60-77 in J.W. Thomas, tech. ed. Wildlife Habitats in Managed Forests: The Blue Mountains of Oregon and Washington. U.S. Dept. Agric. Forest Service, Agriculture Handbook No. 553. Vickery, P.D., M.L. Hunter Jr., and J.V. Wells. 1992. Use of a new reproductive index to evaluate relationships between habitat quality and breeding success. Auk 109:697-705.

128 AND THE WINNER IS… LONGLEAF PINE: RESPONSE OF PLANTED PINES TO VARIOUS CULTURAL TREATMENTS 39 YEARS LATER

April H. Harris (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849) John S. Kush (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849) Ralph S. Meldahl (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849) J. Morgan Varner (Interdisciplinary Ecology Program, Box 118526, University of Florida, Gainesville, FL 32611-8526)

ABSTRACT: Forest managers generally do not seriously consider longleaf pine in their management plans because of the slow early growth compared to slash and loblolly pines. However, many forest managers believe that the long-term growth rate of longleaf pine is superior and it produces a higher quality product. An intensive culture study of longleaf, loblolly, and slash pine was initiated in 1960 on the Harrison Experimental Forest, near Saucier, MS. Five cultural treatments were applied to high specific gravity populations and five to average specific gravity populations for each species. The five cultural treatments were: 1) no cultivation and no fertilizer, 2) cultivation and no fertilizer, 3) cultivation and low levels of fertilizer, 4) cultivation and medium levels of fertilizer, and 5) cultivation and high levels of fertilizer. The study was re-measured during the 1999-2000 winter, the first time since 1985 when the study was 25 years old. The following table compares 1985 data with 1999:

Trees/acre Basal area/acre DBH (inches) Height (feet) Volume (ft3/acre) 1985 1999 1985 1999 1985 1999 1985 1999 1985 1999 Longleaf pine 154a 140a 49.51a 68.51a 7.44b 9.42ab 55.1a 68.7b 2802ab 4918a Slash pine 151a 120b 43.06b 59.90a 7.83a 9.66a 56.7a 71.9a 2908a 4475a Loblolly pine 132b 99c 47.95ab 50.02b 7.47b 9.17b 51.0c 63.4c 2467b 3580b

(Different letters indicate a statistically significant difference between species within a column.)

In addition, each tree was classified as to whether it would make a utility pole or not. Nearly 71% of the longleaf would make a pole while slash and loblolly has only 12.1 and 7.9%, respectively.

INTRODUCTION In the southeastern US, longleaf pine forests may have occupied as much as 90 million acres at the time of European settlement. Intensive exploitation and a lack of regeneration efforts have decreased these once vast forests to fewer than 3 million acres today. Much of these remaining forests are second-growth that survived mostly by chance following logging of the old-growth forests.

Forest managers generally do not seriously consider longleaf pine in their management plans because of the slow early growth compared to slash and loblolly pines. However, many forest managers believe that the long-term growth rate of longleaf pine is superior and it produces a higher quality product. Due to the high quality of longleaf pine and values associated with a fire-maintained longleaf pine ecosystem, many landowners are taking advantage of cost-share programs and recent improvements in longleaf pine regeneration techniques.

An intensive culture study of three southern pines was initiated in 1960 on the Harrison Experimental Forest, near Saucier, MS. With support from the U.S. Forest Service, Southern Research Station, the study was re- measured during the 1999-2000 winter by the Longleaf Pine Stand Dynamics Laboratory located at Auburn University's School of Forestry and Wildlife Sciences. This study will compare differences in growth rates of longleaf, slash, and loblolly pine after 39 years and will compare changes that occurred since the study was last re-measured in 1985.

129 MATERIALS AND METHODS The study is located on the Harrison Experimental Forest, near Saucier, MS, about 25 miles north of Gulfport (see figure above). It had been stocked with second-growth longleaf pines before being clearcut in 1958-59. The soils are well-drained upland, fine sandy loams in the Poarch series and the Saucier-Susquehanna complex. Slope varies from 0 to 8 percent on the gently rolling land.

One-year-old seedlings of longleaf pine, slash pine, and loblolly pine were planted at 10- by 10-foot spacing in 1960. Within each plot, there were 10 subplots, five cultural treatments applied to high specific gravity populations and five to average specific gravity populations. The five cultural treatments were:

U-0 - check, no cultivation and no fertilizer; C-0 - cultivation, but no fertilizer; C-1 - cultivation and a single application of 1,000 pounds of 10-5-5 NPK/acre; C-2 - cultivation and a single application of 2,000 pounds of 10-5-5 NPK/acre; and C-3 - cultivation and a single application of 4,000 pounds of 10-5-5 NPK/acre.

Cultivated plots were cleared of all stumps and slash, then plowed and disked. On check plots, stumps, soil, and competing vegetation were not disturbed. Cultivation consisted of disking 3 times each season for 3 years after planting and mowing in the 4th and 5th seasons. Fertilizer was distributed with an agricultural spreader and disked into the soil in May 1961, 1 year after planting.

The high specific gravity seedlings averaged 0.57 in specific gravity and the average specific gravity seedlings averaged 0.54. The reason for this was considerable concern over possible reduction in wood quality due to increased growth in southern pines because of fertilization and cultivation. Studies had shown that not only did fertilization decrease fiber length and specific gravity, but also trees with higher specific gravity were affected more than trees with lower specific gravity.

Tree heights were measured each year from 1962-1966. Tree height and diameter at breast height (DBH) were measured in 1968 with DBH measured again in 1969 and 1972. Tree height and DBH were last measured in 1985 when the study was 25 years old.

RESULTS Initial statistical analyses were performed across all of the different treatments within species and between species. There were no statistically significant differences among the high and average specific gravity populations. Therefore, for purposes of this presentation the high and average specific gravity populations were averaged and used for the comparisons.

The table presents the statistical comparisons across all of the treatments comparing 1985 to 1999. Mortality among longleaf pine was low compared to the other species, losing only 11 trees/acre compared o 31 and 33 trees/acre for slash and loblolly pine. Basal area increased the most for longleaf while loblolly, due to high mortality had very little basal area growth. Moreover, while longleaf pine had a lower total volume in 1985 compared to slash pine, it averaged 500 ft3/acre more in 1999.

Trees/acre Basal area/acre DBH (inches) Height (feet) Volume (ft3/acre) 1985 1999 1985 1999 1985 1999 1985 1999 1985 1999 Longleaf pine 154a 140a 49.51a 68.51a 7.44b 9.42ab 55.1a 68.7b 2802ab 4918a Slash pine 151a 120b 43.06b 59.90a 7.83a 9.66a 56.7a 71.9a 2908a 4475a Loblolly pine 132b 99c 47.95ab 50.02b 7.47b 9.17b 51.0c 63.4c 2467b 3580b

(Different letters indicate a statistically significant difference between species within a column.)

Volume was calculated with a simple equation of (DBH*DBH)*Height.

130 THE FOLLOWING IS A LIST OF PUBLICATIONS FROM THIS STUDY: Schmidtling, R. C. 1972. Replacement planting with potted southern pines on research plots. U.S. Department of Agriculture, Forest Service, Reseacrh Note SO-146. Schmidtling, R.C. 1973. Intensive culture increases growth without affecting wood quality of young southern pines. Canadian Journal of Forest Research 3:565-573. Schmidtling, R. C. 1984. Early intensive culture affects long-term growth of loblolly pine trees. Forest Science 30(2):491-498. Schmidtling, R.C. 1985. Species and cultural effects on soil chemistry in a southern pine plantation after 24 years. Pages: 573-577, in (Eugene Shoulders, ed.) Proceedings of the third biennial southern silvicultural research conference; 1984 November 7-8; Atlanta, GA: General Technical Report SO-54. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station, 589 p. Schmidtling R.C. 1987. Relative performance of longleaf compared to loblolly and slash pines under different levels of intensive culture. 395-400, in (Douglas R. Phillips, ed.) Proceedings of the fourth biennial southern silvicultural research conference; 1986 November 4-6; Atlanta, GA: General Technical Report SE-42. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southeastern Forest Experiment Station, 598 p. Smith, L. F., and R. C. Schmidtling. 1970. Cultivation and fertilization speed early growth of planted southern pines. U.S. Department of Agriculture, Forest Service, Tree Planter’s Notes 21(1):1-3.

ACKNOWLEDGEMENTS The authors wish to thank Dr. Ron Schmidtling for collecting years of data and for the opportunity to re- measure the study. In addition, we would like to thank to staff of the Harrison Experimental Forest. This study was funded through the generous support of the USDA Forest Service Southern Research Station and especially Charles K. McMahon. Assistance in the field was provided by Richard Sampson, Chadwick Avery, Dennis Shaw, Eric Reynolds, Bryan Lindsey and Lee Helton.

131 THE EFFECT OF THE CONSERVATION RESERVE PROGRAM ON LONGLEAF PINE PLANTING IN GEORGIA

Rick Hatten (Georgia Forestry Commission, P.O. Box 819, Macon, GA 31202-0819)

ABSTRACT: Tree planting is the primary practice implemented in Georgia under the Conservation Reserve Program (CRP). Since the inception of the CRP, loblolly pine and slash pine were the predominate species planted on approximately 800,000 eligible acres in Georgia. Eligibility of acres was determined by a cropping history and the erodibility of the soil. The field had to have grown a crop during two of the previous five years and the soil had to be classified as highly erodible by the Natural Resources Conservation Service.

A Longleaf Pine National Conservation Priority Area (CPA) was established in 1998 due to the efforts of several State and Federal Agencies in the Southern United States and the Longleaf Alliance. Land eligibility changed within the CPA with the soils no longer having to be classified as highly erodible. The land also had to fall within the former natural range of longleaf pine.

The 18th and 20th CRP enrollments for tree planting in Georgia totaled approximately 160,000 acres with over 138,000 acres devoted to longleaf pine. It is unlikely that the 138,000 acres would have qualified for tree planting under CRP without the CPA. There still are approximately 750,000 acres under CRP contract in Georgia with 17.3% in longleaf pine.

132 PROTECTING AND RESTORING LONGLEAF PINE FORESTS ON THE KISATCHIE NATIONAL FOREST IN LOUISIANA

James Haywood (USDA, Forest Service, 2500 Shreveport Highway, Pineville, LA 71360) Michael Elliott-Smith (USDA, Forest Service, 2500 Shreveport Highway, Pineville, LA 71360) Finis Harris (USD Agriculture, Forest Service, 2500 Shreveport Highway, Pineville, LA 71360) Alton Martin (USDA, Forest Service, 2500 Shreveport Highway, Pineville, LA 71360)

ABSTRACT: Longleaf pine (Pinus palustris Mill.) forests once constituted a major ecosystem in the Southern United States stretching from southeastern Virginia south to central Florida and west into East Texas. These forests covered a wide range of site conditions, from wet pine flatwoods to dry mountain slopes. Intensive exploitation reduced the extent of old-growth longleaf forests to 20 million acres by 1935, 12 million by 1955, 3.8 million by 1985, and 3.2 million acres by 1993.

The continued loss of longleaf pine forests has endangered or threatened nearly 200 associated taxa of vascular plants and several vertebrate species. Protecting the remaining longleaf pine forest and restoring longleaf pine plant communities within their historical ranges is paramount in protecting threatened species from extinction.

In January 1993, the Kisatchie National Forest and Southern Research Station began cooperative Ecosystem Management Projects to monitor how prescribed burning affects overstory and midstory trees and shrubs and understory shrub and herbaceous vegetation and to demonstrate how group selection cutting and shelterwood with reserved trees can be used to restore old growth attributes in longleaf pine forests. Our poster outlines this cooperative effort over the last eight years.

INTRODUCTION Longleaf pine (Pinus palustris Mill.) forests once constituted a major ecosystem in the Southern United States stretching from southeastern Virginia south to central Florida and west into East Texas (Outcalt and Sheffield 1996). These forests covered a wide range of site conditions, from wet pine flatwoods to dry mountain slopes. Intensive exploitation reduced the extent of old-growth longleaf forests to 20 million acres by 1935, 12 million by 1955, 3.8 million by 1985, and 3.2 million acres by 1993.

The continued loss of longleaf pine forests has endangered or threatened nearly 200 associated taxa of vascular plants and several vertebrate species (Brockway et al. 1998). Protecting the remaining longleaf pine forest and restoring longleaf pine plant communities within their historical ranges is paramount in protecting these threatened species from extinction.

In January 1993, the Kisatchie National Forest and Southern Research Station began cooperative Ecosystem Management Projects to monitor how prescribed burning affects overstory and midstory trees and shrubs and understory shrub and herbaceous vegetation and to demonstrate how group selection cutting and shelterwood with reserved trees can be used to restore old growth attributes in longleaf pine forests. The desired future condition is open stands of pure longleaf pine, with few if any midstory hardwood trees, over rich and productive herbaceous plant communities intermixed with understory hardwoods kept in check by repeated prescribed burning. Pure longleaf pine is defined as at least 80% of the overstory basal area, stem numbers, or volume being longleaf pine (Helms 1998).

MONITORING SITES Sites were selected from existing stands of predominately longleaf pine that were repeatedly prescribed burned in the past and would be burned again within several months of selection. Sites were selected on the Calcasieu (originally the Vernon and Evangeline), Catahoula, Kisatchie, and Winn Ranger Districts of the Kisatchie National Forest near Leesville, Alexandria, Natchitoches, and Winnfield, Louisiana, respectively. All sites were within the upland longleaf pine forest type of the West Gulf Coastal Plain. The mean January and July temperatures range from 34 and 93 oF across these sites (Louisiana Office of State Climatology 1999). Annual rainfall ranges from 56 to 60 inches and is well distributed throughout the year. The growing season is from 230 to 260 days long.

133 FINDINGS TO DATE In our work, overstory and midstory basal area and canopy cover and number and stature of understory trees and shrubs were inversely related to current-year herbaceous plant production. This was not surprising because the inverse relationship between longleaf pine basal area and herbaceous plant productivity is well known (Grelen and Lohrey 1978). Although overstory basal areas increased in our forests, repeated prescribed burning was able to keep herbaceous vegetation from disappearing because fire reduces understory loblolly pine (P. taeda L.) and hardwood growth and stature over a number of years (Chen and others 1975, Grelen and Epps 1967), and litter does not accumulate sufficiently to smother herbaceous plants.

Nevertheless as a pine canopy closes, the ill effects of shading by the overstory and competition for water and nutrients cannot be entirely overcome by applying a fire regime and herbaceous vegetation inevitably declines (Wolters 1982). Indeed, the decline of herbaceous communities on our repeatedly burned forest sites was unfortunate because the desired future condition may not be reached.

Thinning of the overstory coupled with continued prescribed burning to reduce small woody vegetation improves conditions for herbaceous plant development (Grelen and Lohrey 1978), and thinning may be needed if the desired future condition is to be reached or maintained on our upland longleaf pine sites.

Another finding was the lack of longleaf pine recruitment on our sites. Longleaf pine regeneration cannot develop sufficiently under established forest canopy unless basal areas are reduced to below 30 ft2/acre (Boyer 1993). So some means of stand conversion may be needed to obtain advanced longleaf regeneration.

MANAGEMENT IMPLICATIONS If grass-stage longleaf pine seedlings develop in open ranges of established grass cover, as were common in the West Gulf Region as late as the early 1960s, the seedlings reach sufficient girth to tolerate burning. However, open-range conditions are rare today, and forest managers wishing to restore upland mixed pine or mixed pine-hardwood forests to pure longleaf pine often have stands with poorly developed herbaceous plant communities.

Under today’s conditions, we believe that on most upland sites a series of preharvest treatments are needed to ensure the restoration of longleaf pine and associated plant communities. The treatments would mostly involve prescribed burning and thinning prior to final overstory removal. These treatments are necessary to establish a herbaceous plant community under the existing overstory before final harvest.

Prescribed burning should be the first treatment applied as fire is considered a necessary management tool for preparing sites for longleaf pine regeneration (Wahlenberg 1946). Implementing a timely fire regime is also needed for fuel reduction, and reducing midstory trees and shrubs.

However, to accomplish complete midstory control, herbicides or mechanical means may have to be used where vegetation is too large to control with prescribed burning. Also, a herbicide or mechanical treatment may be required where managers must restore certain plant communities as quickly as possible because the effects of a single prescribed burn are often transitory, and a series of burns over many years must be completed and maintained to have lasting changes in plant communities (Brockway et al. 1998).

As the midstory is controlled and accumulated litter removed, grasses and forbs will naturally reestablish on forested uplands in the West Gulf Region (Haywood et al. 1998). This herbaceous plant cover supports low intensity burns with minimal smoke. The continued implementation of fire is essential and should be applied when woody stems start to become reestablished in the understory. Frequency of burns is dependent upon site productivity and the desired or existing plant community.

Thinning treatments remove the immediate loblolly pine seed source and reduce canopy cover allowing more sunlight to reach the forest floor. This further favors the natural recovery of herbaceous vegetation (Grelen and Lohrey 1978). Once a herbaceous understory is in place, the overstory can be harvested and longleaf pine seedlings planted. Or if longleaf pines are present, a natural regeneration system can be used employing either shelterwood, shelterwood with reserves, or group selection methods. The latter two methods of cut maintain mature trees on the site. This may be the most ecologically beneficial provided the distribution of

134 the mature trees is controlled to favor longleaf pine recruitment in openings (Palik et al. 1997). Brockway and Outcalt (1998) recommend group selection with the openings being 130 to 165 ft in diameter (0.3 to 0.5 ac). This provides enough open space for the intolerant longleaf pine seedlings to develop without intensive intraspecific competition with adult longleaf pines for light, water, and nutrients.

CONCLUSIONS As the stand canopy becomes more open in character with the continual application of fire and thinnings, the recovering grass-dominated herbaceous plant community intermixed with pine needles will provide fuels for prescribed fires to control hardwoods and loblolly pine seedlings and promote the establishment of the longleaf pine regeneration. At this stage of stand development, spring burns will best help establish longleaf pine (Haywood and Grelen 2000, Grelen 1975). Although not all of the desired herbaceous plants may now be found in the stand, over time the desired species should reestablish themselves in the West Gulf Region (Haywood et al. 1998).

LITERATURE CITED Boyer, W.D. 1993. Long-term development of regeneration under longleaf pine seedtree and shelterwood stands. South. J. Appl. For. 17:10-15. Brockway, D.G. and K.W. Outcalt. 1998. Gap-phase regeneration in longleaf pine wiregrass ecosystems. For. Ecol. Manage. 106:125-139. Brockway, D.G., K.W. Outcalt, and R.N. Wilkins. 1998. Restoring longleaf pine wiregrass ecosystems: plant cover, diversity and biomass following low-rate hexazinone application on Florida sandhills. For. Ecol. Manage. 103:159-175. Chen, M.Y., E.J. Hodgkins, and W.J. Watson. 1975. Prescribed burning for improving pine production and wildlife habitat in the hilly coastal plain of Alabama. Bull. 473. Auburn, AL: Auburn University, Alabama Agricultural Experiment Station. 19 p. Grelen, H.E. 1975. Vegetative response to twelve years of seasonal burning on a Louisiana longleaf pine site. Res. Note SO-192. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station. 4 p. Grelen, H.E. and E.A. Epps, Jr. 1967. Season of burning affects herbage quality and yield on pine-bluestem range. J. Range Manage. 20:403-404. Grelen, H.E. and R.E. Lohrey. 1978. Herbage yield related to basal area and rainfall in a thinned longleaf plantation. Res. Note SO-232. New Orleans, LA: U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station. 4 p. Haywood, J.D., A. Martin, Jr., H.A. Pearson, and H.E. Grelen. 1998. Seasonal biennial burning and woody plant control influence native vegetation in loblolly pine stands. Res. Paper. SRS-14. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 8 p. Haywood, J.D. and H.E. Grelen. 2000. Twenty years of prescribed burning influence the development of direct-seeded longleaf pine on a wet pine site in Louisiana. South. J. Appl. For. 24(2):86-92. Helms, J.E. editor. 1998. The dictionary of forestry. Washington, DC: Society of American Foresters. 210 p. Louisiana Office of State Climatology. 1999. Louisiana monthly climate review. Baton Rouge, LA: Louisiana State University, Southern Regional Climate Center. Vol 19. 98 p. Outcalt, K.W. and R.M. Sheffield. 1996. The longleaf pine forest: trends and current conditions. Resource Bull. SRS-9. Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 23 p. Palik, B.J., R.J. Mitchell, G. Houseal, and N. Pederson. 1997. Effects of canopy structure on resource availability and seedling responses in a longleaf pine ecosystem. Can. J. For. Res. 27:1458-1464. Wahlenberg, W.G. 1946. Longleaf pine its use, ecology, regeneration, protection, growth, and management. Charles Lathrop Pack Forestry Foundation and USDA For. Serv., Washington, DC. 429 p. Wolters, G.L. 1982. Longleaf and slash pine decreases herbage production and alters herbage composition. J. Range Manage. 35: 761-763.

135 COMPARISON OF ARTHROPODS ON THE BOLES OF LOBLOLLY AND LONGLEAF PINES (PINACEAE)

G. Scott Horn (U.S. Forest Service, Southern Research Station, Forestry Sciences Laboratory, 320 Green St, Athens, GA 30602) James L. Hanula (U.S. Forest Service, Southern Research Station, Forestry Sciences Laboratory, 320 Green St, Athens, GA 30602)

ABSTRACT: Red-cockaded woodpeckers (Picoides borealis) (RCW) forage on the bole of most southern pines. We uses knockdown insecticides to sample the standing crop of arthropods on longleaf (Pinus palustris Mill) and loblolly pine (P.taeda L.), two common pines within foraging habitats of RCW to determine which harbored the greater abundance of potential prey. Longleaf pine had higher arthropod abundance (278 + 44.4/tree) compared to loblolly pine (132 + 132/tree). Oven-dried biomass estimate were also higher on longleaf pine (945 mg + 145/tree) compared to loblolly pine (395 mg + 28/tree). Certain groups were found in significantly higher numbers on longleaf pine, including Thysanura, Hemiptera, and Pseudoscorpiones. The biomass of Blattaria was also much higher on longleaf boles, suggesting that larger arthropods may prefer the bark structure of longleaf pine. We altered the bark surface of longleaf pine to determine if it was bark structure influencing greater numbers of arthropods residing on the tree bole. When the loose bark was removed, we recovered far few arthropods than from unscraped control trees. We also lightly scraped the outer bark of both tree species and found that longleaf pine had significantly more loose, flaking, bark scales than loblolly. These results suggest that it is the bark structure and not the chemical nature of the bark that results in more and larger arthropods on longleaf pine. Retaining or restoring longleaf pine in RCW habitats should increase arthropod availability for this endangered bird and other bark-foraging species.

136 THE ROLE OF MISSISSIPPI LONGLEAF PINE IN RESTORING THE USS CONSTITUTION

H. Glenn Hughes (Mississippi State University Extension Service, Hattiesburg, MS 39401)

ABSTRACT: The USS Constitution, known as “Old Ironsides” is the oldest US Warship still in commission. Her Georgia live oak keel was laid in 1795, and she celebrated her 200th birthday on October 21, 1997. Restored in 1996, she has 1 acre of canvas sails and 8 miles of rigging.

In the early 1900’s, she was almost scrapped, and in the 1920’s, she was restored. Part of the restoration effort involved new decking for the Constitution. The new decking of longleaf pine was cut at the Major- Sowers Sawmill in Tallahala, Mississippi, in July of 1928. She was recommissioned in 1931, and set out undertow for a tour of 90 port cities along the Atlantic, Gulf and Pacific coasts of the United States.

This poster illustrates the historical significance of longleaf pine from southeast Mississippi to the nation’s oldest commissioned warship. The poster shows several black-and-white photographs of the Major-Sowers logging operation in 1928. Included are longleaf pine logs loaded on rail cars pulled by a 50-ton Heisler locomotive, cut decking at the sawmill and loaded on rail cars for transportation, and a general photograph of the Major-Sowers sawmill.

137 TRACKING RARE PLANT SITES AND HABITAT USING GLOBAL POSITIONING AND GEOGRAPHIC INFORMATION SYSTEMS ON THE KISATCHIE NATIONAL FOREST

Philip E. Hyatt (USDA Forest Service, 2500 Shreveport Hwy, Pineville, LA 71360)

ABSTRACT: The Kisatchie National Forest manages about 600,000 acres in central and northern Louisiana. Within that landscape, the Forest tracks 77 rare plant species. We have no known sites of federally listed threatened or endangered plant species. We do track 19 plants that are rare throughout their range "Sensitive" as well as other species that may be common elsewhere, but are rare in Louisiana ("Conservation" species). These plants occur in several habitats, including some found within Longleaf ecosystems: sandywoodlands, bottomland forests, pitcher plant bogs and baygalls, calcareous prairies, upland longleaf pine forests, and other communities. GIS provides a tool for tracking rare plant sites. Polygons outline the extent of communities. Within these "planthab" polygons, points mark specific sites where one or more plants have been found. GPS technology will allow future botanists to relocate these plant sites once they are GPSed. By plugging in coordinates, botanists will be able to return to within 5-10 feet of any previously documented site or track the expansion or contraction of bogs, prairies, and other communities.

138 FLORISTIC STUDY OF THE KISATCHIE NATIONAL FOREST, LOUISIANA, USA

Philip E. Hyatt (USDA Forest Service, 2500 Shreveport Hwy, Pineville, LA 71360)

ABSTRACT: A comprehensive study of the Kisatchie National Forest's flora has started. Researchers have published a variety of papers in the past on parts of the flora of the Kisatchie National Forest in central and northern Louisiana, U. S. A. Bog Research (Michael and Barbara MacRoberts) has published a variety of reports on the flora of specialists habitats, especially bogs, glades, and prairies. Range researchers from the late 1930s to the mid 1980s created a small 5000 specimen herbarium (SFRP) concentrating primarily on grasses and other associated species found in range allotments. The author began collecting Carex sedges and doing some general collecting in 1995, and has prepared a database of SFRP holdings. Data is currently available at the U. S. D. A. PLANTS database, and will soon be available on the Internet on the Kisatchie National Forest website. Ongoing work lays a course that will continue general surveys for the Kisatchie's flora, and document that flora by Ranger District and sub-unit. Emphasis currently centers on documenting rare plant sites, the author's area of interest (Cyperaceae), and general floristics. SFRP has an active herbarium program with emphasis on donation to other herbaria rather than trade due to limited space for an extensive collection.

139 LONGLEAF PINE ECOSYSTEM RESTORATION ON THE NATIONAL FORESTS AND SAVANNAH RIVER INSTITUTE

R.M. Jeffers (USDA Forest Service, Southern Region, Atlanta, GA) K.E. Stoneking (USDA Forest Service, Southern Region, Atlanta, GA) D.J. Tomczak (USDA Forest Service, Southern Region, Atlanta, GA)

The National Forests (NF) and Savannah River Natural Resource Management and Research Institute (SRI), Department of Energy, located in the 13 states of the Southern Region, USDA Forest Service, include more than 12.5 million acres of land included in the area extending from Virginia, south to Florida, and west to east Texas. National Forests and the Savannah River Institute occur in eight states within the historical range of the longleaf pine ecosystem including North Carolina, South Carolina, Georgia, Florida, Alabama, Mississippi, Louisiana, and Texas. The National Forest System’s continuous inventory of stand conditions (CISC) database showed that acres of longleaf pine forest cover type and mixtures of longleaf pine and other pine and hardwood species existed on 668 thousand acres of NF and SRI lands in 1988. The 2000 CISC database shows that longleaf pine and longleaf pine-other species mixtures now occur on 831 thousand acres of these lands, an increase of 31 percent compared to 1988. Current Forest Land and Resource Management Plans (Forest Plans) for the NF and SRI call for a total of 1.245 million acres of desired future condition acres (DFC) of longleaf pine types (Figure 1 and Table 1). This amounts to a doubling of longleaf acres compared to 1988.

The NF and SRI are making restoration of the longleaf pine ecosystem a high priority issue as their Forest Plans are being revised. Forest Plans are required by the Forest and Rangeland Renewable Resources Planning Act of 1974, as amended by the National Forest Management Act of 1976. Forest Plan revision occurs when conditions have changed significantly, or at least every 15 years. A Forest Plan does not direct specific management activities for specific locations. The Forest Plan does significantly influence the design, execution, and monitoring of site-specific forest activities. Environmental documents, which accompany the Forest Plan, provide data disclosing the environmental consequences of proposed management alternatives. The Plan also discloses the effects of proposed alternatives and how they respond to public issues and concerns. Revised Forest Plans for the NF having significant acres of longleaf pine and the SRI in the Southern Region, USDA Forest Service call for continuing increases in longleaf pine acres, control of undesirable species and other competing vegetation in the mid-story, and improved health of the understory plant communities by mechanical and chemical means, and through the use of periodic growing and dormant season prescribed fires (McMahon, Tomczak, and Jeffers 1997).

The longleaf pine ecosystem will not be restored to the full range of sites and conditions, which existed prior to pre-European settlement. Longleaf pine ecosystem restoration on the NF and SRI will be a slow and continuing process. However, it will be restored, maintained and enhanced to the fullest extent possible; and the restored longleaf pine ecosystem on the NF and SRI lands will constitute about one-third of all longleaf pine acres.

LITERATURE CITED McMahon, C.K., D.J. Tomczak, and R.M. Jeffers. 1997. Longleaf pine ecosystem restoration: The role of the USDA Forest Service. In: Kush, John S., comp. Proceedings of the Longleaf Pine Ecosystem Restoration Symposium, presented at the Society of Ecological Restoration 9th Ann. Intern. Conf. – Ecological Restoration and Regional Strategies, November 12-15, 1997. Longleaf Alliance Rep. No. 3. Fort Lauderdale, FL. 20-31.

140 Figure 1 - Acres of longleaf pine on the National Forests and Savannah River Institute 1/

Thousand Acres 300

250

200

150

100

50

0 TX LA MS AL FL SRI SC NC

1988 2000 DFC

1/ Includes longleaf pine and mixtures of longleaf and other species Source: R8 database CISC

141

Table 1 - Acres of longleaf pine on the National Forests and Savannah River Institute 1/

Forest 1988 2000 DFC

North Carolina NF's in N. Carolina 12,208 15,052 19,403

South Carolina Francis Marion & Sumter NF's 32,658 48,475 54,555 Savannah River Institute 33,913 43,320 86,089

Georgia Chattahoochee NF 0 0 300

Florida NF's in Florida 175,994 209,441 221,121

Alabama NF's in Alabama 128,575 153,392 270,619

Mississippi NF's in Mississippi 182,283 205,695 275,029

Louisiana Kisatchie NF 81,526 127,314 210,819

Texas NF's in Texas 20,805 28,417 107,225

Total 667,962 831,106 1,245,160

Increase (%) since 1988 31.1 96.4

1/ Includes longleaf pine and mixtures of longleaf pine and other species.

142 EARLY GROWTH RESPONSE OF LONGLEAF PINE (PINUS PALUSTRIS MILL.) SEEDLINGS TO LIGHT AVAILABILITY

Shibu Jose (School of Forest Resources and Conservation and West Florida Research and Education Center, University of Florida, PO Box 3634, Milton, FL 32572) Sara Merritt (School of Forest Resources and Conservation and West Florida Research and Education Center, University of Florida, PO Box 3634, Milton, FL 32572)

ABSTRACT: Although there have been studies examining water and nutrients as determinants of longleaf pine seedling growth, little emphasis has been placed on light availability. Hence, the objectives of the present study were to (1) examine the effects of light on the growth of longleaf pine seedlings, (2) determine the interactive effects of light and soil resource (water and nitrogen) availability on longleaf pine growth, and (3) explore the physiological basis for the observed growth differences. The experiment was conducted in a greenhouse with one- year old longleaf pine seedlings grown in 11.3 L pots. The experimental design was a split-plot factorial with high and low levels of light, water, and nitrogen as the factors. The eight treatment combinations were replicated 15 times in a split plot design. Data were collected on survival, root collar diameter, and height on a monthly basis. Physiological data (net photosynthesis and transpiration) were also collected every month. First year results indicate that seedlings grown under high levels of light (800-900 µmol m-2 -s) had higher diameter (31%) and height (32%) growth compared to seedlings grown under low levels of light (300-350 µmol m-2 -s ). Net photosynthesis for seedlings grown under high light was 59% higher than for those grown under low light. The results indicate that competition for light can also play a major role in determining the growth rate of longleaf pine seedlings.

INTRODUCTION Longleaf pine (Pinus palustris Mill.) forests were one of the most extensive ecosystems in North America prior to European settlement (Landers et al., 1995). The presettlement forests in the South contained more than 74 million acres of longleaf pine-dominated stands growing on a range of site and soil conditions (Frost, 1993). However, with the clearing of land for agriculture and logging operations without adequate regeneration efforts, the area under longleaf pine decreased considerably. For example, 27 % of the longleaf pine land had been converted to farmland by 1900 (Frost, 1993). Conversion of land into fields of more aggressive pine species such as loblolly pine (Pinus taeda L.) and slash pine (Pinus elliottii Engelm.) and exclusion of fire from the landscape have resulted in further decrease in longleaf pine acreage during the last several decades. Most recent estimates show that only 4 % of the original area remain today (Brockway and Outcalt, 1998) making it one among the most threatened ecosystems in North America. The fact that most of the remaining longleaf pine stands are aging without adequate regeneration and replacement (Kelly and Bechtold, 1990) poses a serious threat to the sustainability of these unique forests.

Restoration of longleaf pine forests, including afforestation of old-fields with longleaf pine, has risen to the top of conservation efforts in the southeastern United States in recent years. Successful restoration efforts can result in new stands of longleaf pine, providing a wide array of ecological, social, and economic benefits. However, many private landowners and public land managers are concerned with the long residence time of longleaf pine in the grass stage. Previous studies have examined the role of water and nutrients in controlling longleaf pine seedling growth and establishment. However, little emphasis has been placed on the role of light availability. Given the fact that millions of longleaf pine seedlings are planted every year in canopy gaps it is important to investigate the effects of light availability on longleaf pine establishment and growth. Hence, the objectives of the present study were to (1) examine the effects of light on the growth of longleaf pine seedlings, (2) determine the interactive effects of light and soil resource (water and nitrogen) availability on longleaf pine growth, and (3) explore the physiological basis for the observed growth differences.

METHODS The experiment was conducted in a greenhouse at the University of Florida-Milton campus in Milton, FL. One-year-old nursery-grown longleaf pine seedlings of uniform size from a wild seed source were planted, one per pot, in 11.3 L plastic pots filled with a planting medium of coarse sand and pine bark of very low fertility, similar to the soils of the lower coastal plain. The experimental design was a split-plot factorial with light (high and low), nitrogen (high and low) and water (high and low) as the factors. The eight treatment

143 combinations were replicated fifteen times in a split-plot design. The light treatment was the main plot factor with factorial combinations of nitrogen and water applied randomly to the pots. Shade cloth was used to create the low light treatment where seedlings received a photosynthetically active radiation (PAR) of 300 to 350 µmol m-2 –s. Seedlings grown under high light treatment received a PAR of 800 to 900 µmol m-2 –s. Nitrogen treatments consisted of either 40 or 400 kg ha-1 yr-1 of ammonium nitrate applied at four month intervals. Other nutrients were kept at nonlimiting levels in all the pots. The high water treatment involved watering seedlings five days a week. Seedlings were watered only once a week (in the beginning) or once in two weeks (after establishment) in the low water treatment. Seedling root collar diameter and height were measured every month for one year. Net photosynthesis on three seedlings per treatment was also measured on a monthly basis using an infrared gas analyzer.

RESULTS AND DISCUSSION Longleaf pine seedling had a higher growth rate in the high light treatment than the low light treatment. Averaged across all nitrogen and water treatments, seedlings grown under high light treatment exhibited 31%

16 6 14 Low Light Low Light High Light 5 High Light 12 10 4 8 3

6 2 4 1 2 0 Height(cm growth per year) 0 Diameter growth (mm per year) (mm per growth Diameter Low Nitrogen High Nitrogen Low Nitrogen High Nitrogen

16 6 Low Light Low Light 14 5 High Light High Light 12 4 10 8 3 6 2 4 1 2 Height growth (cm per year) per (cm growth Height

Diameter growth (mm per year) 0 0 Low Water High Water Low Water High Water higher diameter growth and 32% higher height growth compared to seedlings grown under low light. High nitrogen and water increased seedling height and diameter growth substantially under high light treatment.

Figure 1. Longleaf pine seedling diameter and height growth as influenced by light availability

16 16 14 Low Light 14 Low Light High Light High Light 12 12 10 10 8 8 6 6 4 4 2 2 0 0 Low Nitrogen High Nitrogen Low Water High Water NetPhotosynthesis (UmolCO2 /s) / m2

Figure 2. Longleaf pine net photosynthesis as influenced by light availability.

144 Measurements were made in mid June but had little or no effect in the shade treatment (Figure 1). Net photosynthesis was also higher for seedlings grown under high light environment. For example, averaged across all treatments, a 59% increase in net photosynthesis was observed for seedlings under the high light treatment compared to the shade treatment. While increased nitrogen had no effect on net photosynthesis in the low light treatment, higher light level increased net photosynthetic rate considerably (Figure 2). Water stress, however, reduced net photosynthesis in longleaf seedlings. Despite increased light level, seedlings grown under low water treatment exhibited only a slight increase (26%) in net photosynthesis whereas seedlings in the high water treatment showed a greater gain (78%).

Survival and growth of seedlings in relation to light availability have been studied extensively in natural environments (Canham, 1988; Sipe and Bazzaz, 1995; Walters and Reich, 1995; Chen and Klinka, 1998). Increased growth as a result of increased light availability has been demonstrated in several hardwoods and conifers. In natural forests, increased light on the forest floor is often associated with increased gap size. For example, the intensity of light has been shown to increase from 8% to 75% of full sun when gap size increases from 2 m2 to 220 m2 (Jackson, 1959). Light availability on the forest floor is generally higher in longleaf pine forests as a result of an open savanna-like structure. However, modest increase in light levels and associated increase in seedling growth have been reported in response to canopy disturbances that create larger gaps in longleaf forests (Palik et al., 1997). Our results indicate that light availability can be a major factor in determining the residence time of longleaf seedlings in the grass stage. Silvicultural prescriptions should also aim at alleviating the competition for light between longleaf pine seedlings and competing overstory and understory vegetation so that residence time in the grass stage can be reduced.

REFERENCES Brockway, D. G. and Outcalt, K.W. 1998. Gap-phase regeneration in longleaf pine wiregrass ecosystems. For. Ecol. Manage. 106:125-139 Canham, C.D. 1988. Growth and canopy architecture of shade tolerant trees: response to canopy gaps. Ecology 69:786-795. Chen, H.Y.H., Klinka, K. 1998. Survival, growth, and allometry of planted Larix occidentalis seedlings in relations to light availability. Forest Ecology and Management. 106:169-179. Frost, C.C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. In: Herman, S.M. (ed.) Proceedings of the tall Timbers Fire Ecology Conference. Tall Timbers Research Station, Tallahassee, FL, pp. 17-43. Jackson, L.W.R. 1959. Relation of pine overstory opening diameter to growth of pine production. Ecology, 40:478-480. Kelly, J.F., Bechtold, W.A. 1990. The longleaf pine resource. In: Farrar, R.M. (ed.), Proceedings of the symposium on the management of longleaf pine. USDA For. Service Southern Experiment Station General Technical Report SO-75. New Orleans, LA pp. 11-22. Palik, B.J., Mitchell, R.J., Houseal, G., Pederson, N. 1997. Effects of canopy structure on resource availability and seedling responses in a longleaf pine ecosystem. Can. J. For. Res. 27:1458-1464. Sipe, T.W. and Bazzaz, F.A. 1995. Gap partitioning among maples in central New England: survival and growth. Ecology 76:1587-1602. Walters, M.B., Reich, P.B. 1995. Are shade tolerance, survival, and growth linked? Low light and nitrogen effects on hardwood seedlings. Ecology, 77:841-853.

145 LONGLEAF PINE CONSERVATION IN THE CAROLINAS: OVERCOMING BARRIERS TO PRESCRIBED BURNING WITH POLICY CHANGE

Mark Knott (Curriculum in Ecology, The University of North Carolina at Chapel Hill.Current address: 611 NW 14th Street, Corvallis, OR 97330)

ABSTRACT: The occurrence of frequent fire is necessary to maintain and enhance the quality of existing longleaf pine communities. However, expanding the use of prescribed fire to restore longleaf pine habitat in the Southeastern United States remains a formidable challenge. Legal constraints, narrow burn windows, residential development, lack of qualified personnel and other factors may diminish the willingness of land managers to use prescribed fire as frequently as their management objectives might require. An assessment of how these factors influence land managers’ perception of risk could provide direction for policies designed to increase the use of prescribed burning, which could, in turn, lead to better longleaf management.

I am conducting a written survey of public and private land managers who manage at least 1000 acres of longleaf pine in the Carolinas. The research design and survey examine management objectives (economic return, endangered species, etc.), use of prescribed fire (frequency, extent, and season), perceived barriers to the use of prescribed fire, and the impact of specific government policies on the use of prescribed burning (the recently passed “certified burner laws” of North and South Carolina and the Endangered Species Act). Finally, the survey prompts respondents to assess potential incentive programs that may encourage them to expand their use of prescribed fire.

I expect to find that land managers do not use fire as frequently as they might , because state and federal laws do not adequately reduce barriers to prescribed burning. Also, I expect the impact of legal barriers to be more significant than administrative, economic, social, and technical factors.

To better understand the effects of the social, legal, economic, administrative, and technical barriers that limit prescribed burning in longleaf pine ecosystems, I developed a survey following the Tailored Design Method (Dillman 2000). The survey addresses my three research objectives: 1) quantify prescribed burning activity by managers of ≥ 1000 acres of longleaf pine in the Carolinas, 2) determine the perceived significance of factors that limit prescribed burning activity on both private and public longleaf lands in the Carolinas, and 3) gauge the capacity of new policies to promote prescribed fire usage in longleaf ecosystems.

INTRODUCTION At the time of European settlement, a bi-layered forest characterized by an herbaceous understory with an open longleaf pine (Pinus palustris Mill.) overstory dominated some 92 million acres of the southeastern United States (Frost 1993). Low-intensity surface fires swept across the landscape every 2 to 8 years renewing and restoring the forest (Abrahamson and Hartnett 1990, Ware et al. 1993). These fires supported the understory component of herbs and longleaf seedlings by controlling the spread of highly competitive broadleaved species (Christensen 1993a). Within a few years of the fire’s absence the invasion of hardwood, shrub, and mesophytic pine species will eliminate longleaf regeneration on all but the driest sites and herbaceous species diversity will drastically decline (Clewell 1989, Frost 1993, Gilliam and Platt 1999). Thus, longleaf pine ecosystem survival is dependent upon the occurrence of frequent, low-intensity fires.

Following four centuries of land use change longleaf pine ecosystems remnants have been severely degraded through alteration of the landscape and its fire regime. The once massive and continuous longleaf pine forest has been fragmented, planted in off-site pines, tilled up for agriculture, and cleared for urban development (Frost 1993). Of the mere 2.95 million acres of longleaf that remains in the South (Outcalt and Sheffield 1996), a more than 97% reduction of original area, only 0.7% (674,000 acres) truly resembles the original ecosystem (Frost 1993). The latter figure serves as testament to the pervasiveness of fire exclusion throughout the South. Longleaf ecosystem remnants, which were once part of large fire compartments, are now islands in a sea of roads, houses, and fields (Ware et al. 1993), no longer capable of converting lightning strikes into region-wide surface fires (Platt et al. 1988). And so without active prescribed burning, fire will be excluded from these now rare and biologically invaluable ecosystems (Brennan et al. 1998, Johnson and Gjerstad 1998).

146 To meet a variety of objectives many private and public land managers employ prescribed fire as a forest management tool, but by using it they put themselves at potentially serious social, economic, and legal risk (Cleaves and Haines 1995, Cleaves et al. 2000, Haines and Cleaves 1999, Wade and Outcalt 1999). If a fire escapes endangering nearby residences, shifting winds blow smoke onto a nearby roadway or smoke sensitive structure, or local residents are forced to stay indoors because of temporarily poor air quality, land managers and public agencies run the risk of losing public support for burning efforts and/or being sued for damages (Haines and Cleaves 1999, Wade and Outcalt 1999). These risks and other limiting factors may reduce the willingness of longleaf managers, both public and private, to use prescribed burning (Cleaves and Haines 1995, Cleaves et al. 2000, Haines and Cleaves 1999).

The state legislatures of North and South Carolina have recently come to recognize these risks and the importance of fire as a management tool as well as its benefits to public safety, forest and wildlife resources, the environment, and state economies (North Carolina Prescribed Burning Act, 1999). Each state has passed laws designed to “encourage prescribed burning in forests by forest landowners” (North Carolina Prescribed Burning Act 1999) by protecting those willing to be certified by the state. As long as these “certified burners” follow specific standards while burning, their legal liability is partially reduced, which hypothetically will encourage the responsible use of prescribed fire (see South Carolina Prescribed Fire Act, 1994 and North Carolina Prescribed Burning Act, 1999).

While these Acts may give confidence to some forest landowners who use prescribed fire, protecting them from legal liability is only a part of encouraging prescribed burning (Bonnie 1997, Cleaves and Haines 1995, Cleaves et al. 2000, Haines and Cleaves 1999, Wade and Outcalt 1999). In the National Forest System, forest managers may have their burn programs restricted by air quality and smoke management regulations, narrow burn windows, funding problems, and agency risk reduction policies (Cleaves et al. 2000). Industrial forest landowners may burn less because of concerns about site productivity and tree growth, insurance costs (Haines and Cleaves 1999), and the difficulties of complying with internal, state, and federal policies that regulate burning (Wade and Outcalt 1999). Non-industrial private forest landowners may limit their burning due to increasing costs of burn crews and insurance, increased regulation (Cleaves and Haines 1995), and fear of restriction from the Endangered Species Act (Bonnie 1997, Costa 1995a, Costa 1995b). The many barriers not addressed by the “certified burner” laws of North and South Carolina suggest that policy makers may not be fully cognizant of the factors that limit prescribed burning, which, in turn, limits its ecological and social benefits.

Direct conflicts between federal environmental acts may also hamper the use of prescribed fire. The ecosystem management efforts of federal agencies (Estill 1999) and §7(a)(1) of the Endangered Species Act explicitly promote the use of prescribed fire to aid longleaf ecosystem management efforts and endangered species recovery, respectively. However, other federal legislation directly compromises these goals. The Clean Air Act regulates, as one of six criteria pollutants, particulate matter, the primary pollutant from the burning of vegetation. As a result, regulation of prescribed burning to meet national and regional air quality goals, may adversely affect prescribed burning programs (Hauenstein and Siegel 1980), ecosystem management, endangered species recovery, and longleaf ecosystem survival. Exploring the importance and impact of these factors on prescribed burning in longleaf pine ecosystems is the focus of this research.

METHODS To explore 1) prescribed burning activity by the managers of privately and publicly owned longleaf ecosystems, 2) the impact of factors that may limit prescribed burning therein, and 3) the potential effectiveness of policies designed to encourage prescribed burning, a self-administered survey was mailed to 85 individuals (25 of public lands, 6 of forest industry, and 54 of non-industrial private forest lands) who each manage ≥ 1000 acres of longleaf in the Carolinas. Respondents characterized their management objectives, use of prescribed fire, and the effects of social, legal, economic, administrative, and technical barriers on longleaf management. Similarly, they evaluated particular policy changes that might promote active burning programs in longleaf ecosystems.

Full analysis of the survey will also include the following questions: 1) How prevalent are growing season burns? 2) Which factors promote the use of growing season burns? 3) Which factors discourage the use of growing season burns by longleaf managers? 4) Which factors most influence the cost of burning? 5) How

147 will prescribed fire activity change over the next 10 years? 6) How effective are Safe Harbor Agreements and “certified burner” laws?

RESULTS 1. Prescribed burning activity: In total, 66 respondents (78% response rate) who manage 551,000 acres of longleaf (88% of total 624,500 acres remaining in the Carolinas (Outcalt and Sheffield 1996)) returned completed surveys. 74% burn their longleaf sites on a 1 to 3 year rotation, while 95% burn on at least a 4 to 6 year rotation. These frequencies may closely represent the natural fire return intervals of many longleaf ecosystems (Abrahamson and Hartnett 1990, Christensen 1993b, Glitzenstein 1995). However, land managers also indicated that they were unable to burn more than 225,000 acres of their longleaf sites as frequently as they would like.

2. Limiting factors: Respondents from both public and private burn programs specified four common barriers to burning: 1) smoke management and air quality concerns, 2) liability risk, 3) uncertain weather conditions, and 4) proximity to residential development. Similarly, respondents from both groups were more concerned about the risks presented by nearby residential development to the future of their longleaf site(s) than other potential risk factors (endangered species, estate taxes, fire suppression, etc.). Public land managers also expressed concern for personnel shortages, while private land managers targeted local ordinances.

3. Recommended policy changes: I asked respondents to rate the potential effectiveness of 13 policies that could encourage the expansion of burn programs. Both private and public land managers agreed that: 1) eased particulate matter standards and 2) greater liability protection would help them to more easily meet their management objectives with prescribed fire. Managers of private longleaf lands suggested that restructuring the Endangered Species Act would encourage them to expand their burn programs. They favored two strategies: 1) provision of tax incentives for those who manage for endangered species and 2) greater land use assurance when endangered species are present on private lands.

DISCUSSION 1. Prescribed burning activity: The great majority of respondents (95%) burn their longleaf sites on at least a 4 to 6 year rotation and 75% burn on a 1 to 3 year rotation. These burning rotations closely resemble natural fire return intervals (Abrahamson and Hartnett 1990, Christensen 1993a, Glitzenstein 1995), which should be a positive sign for longleaf ecosystem conservation and restoration. However, Frost (1993) indicated that 74% of the longleaf forest remaining in North Carolina is fire suppressed and heavily invaded by hardwoods and loblolly pine. This discrepancy suggests that survey respondents are only able to burn portions of each longleaf site as frequently as every 1 to 3 or 4 to 6 years.

Respondents also answered the question “Considering the total area of the longleaf site(s) that you manage, approximately what percentage are you NOT able to burn as frequently as you would like?” by marking one of two values (0%, 100%) or one of 4 ranges (1-25%, 26-50%, 51-75%, 76-99%). By multiplying either the exact values (0, 1.00) or the midpoint of the range each respondent marked (0.13, 0.38, 0.63, 0.875) by the area of longleaf each respondent manages, I determined that given less restriction or more incentive, land managers would voluntarily expand burn programs on 36% (ca. 225,000 acres) of the longleaf remaining in the Carolinas (Outcalt and Sheffield 1996). Expanding burn programs to include these acres is critical to the recovery and rehabilitation of longleaf pine ecosystems throughout the Carolinas.

2. Limiting factors: Respondents who manage both public and private lands indicated that 1) smoke management and air quality concerns, 2) liability risk, 3) uncertain weather conditions, and 4) proximity to residential development most limited their willingness to use prescribed burning. In another question, they indicted “residential development” as the biggest threat to the future of their longleaf sites. Clearly these factors are interrelated. Proximity to residential development increases the difficulty of managing smoke, narrows the window of appropriate weather conditions, and makes the specter of litigation more apparent. These challenges will only become more problematic as the human population expands throughout the range of longleaf the Carolinas (Cordell et al. 1998). Thus, more effective management of the wildland-urban interface is crucial to make effective longleaf ecosystem management policy.

148 3. Recommended policy changes: Reducing the barriers to prescribed burning is imperative to the recovery and maintenance of high-quality, biologically rich longleaf communities. Public and private land managers agreed that the two most effective means of expanding prescribed fire usage are: 1) easing particulate matter standards (from the Clean Air Act) for those who burn and 2) providing greater liability protection for certified burners. These measures, if taken by federal and state governments, would give land managers: 1) consistent environmental policy and 2) an enhanced ability to effectively manage natural resources in increasingly developed landscapes.

Private land managers indicated that 1) tax incentives for endangered species management and 2) tax deductions for prescribed burning expenses would encourage them to expand their burn programs. Changes in the tax code could 1) turn the presence of endangered species on private lands into an asset rather than a liability and 2) show that governments are committed to sound land management practices.

CONCLUSIONS To more effectively “encourage prescribed burning in forests by forest landowners” local, state, and federal coordination is required to make meaningful policy change. Local jurisdictions, which have nearly exclusive control over land use on the land in the Carolinas not in federal ownership (Crist et al. 2000, Theobald et al. 2000), could help by steering development away from rural areas or instituting rural building standards (Fried et al. 1999). These changes could more effectively control expansion of the wildland-urban interface, which may benefit longleaf restoration efforts. State governments could act by providing more comprehensive liability protection, providing technical assistance, developing incentive programs, and helping county governments develop land use plans that protect significant natural areas.

The federal government also plays an integral role in longleaf pine ecosystem restoration. By allowing National Ambient Air Quality Standards to accommodate prescribed burning, the federal government could give state governments the authority to ease particulate matter standards for those who conduct prescribed burns for forest and ecosystem management. Tax incentives for those private landowners who manage for endangered species may turn a perceived liability into an asset (Bonnie 1999). An Endangered Species Act with provisions that relieve land use restrictions for those who agree to follow specific management practices may also help to expand prescribed burn programs and longleaf recovery on private lands.

REFERENCES Abrahamson, W. G., and D. Hartnett. 1990. Pine flatwoods and dry prairies. Pages 103-149 in R. Myers and J. Ewel, eds. Ecosystems of Florida. University of Central Florida Press, Orlando, FL. Bonnie, R. 1997. Strategies for conservation of the endangered red-cockaded woodpecker on private lands. Endangered Species UPDATE 14: 45-47. Bonnie, R. 1999. Endangered species mitigation banking: promoting recovery through habitat conservation planning under the Endangered Species Act. The Science of the Total Environment 240: 11-19. Brennan, L. A., R. T. Engstrom, W. E. Palmer, S. M. Hermann, G. A. Hurst, L. W. Burger, and C. L. Hardy. 1998. Whither wildlife without fire? Pages 402-414 in K. G. Wadsworth, ed. The Sixty-third North American Wildlife and Natural Resources Conference. Wildlife Management Institute, Orlando, Florida. Christensen, N. L. 1993a. The effects of fire on nutrient cycles in longleaf pine ecosystems. Pages 205-214 in S. M. Hermann, ed. The Longleaf Pine Ecosystem: Ecology, Restoration, and Management. Tall Timbers Research, Inc., Tallahassee, Florida. Christensen, N. L. 1993b. Fire regimes and ecosystem dynamics. Pages 233-244 in P. J. Crutzen and J. G. Goldammer, eds. Fire in the Environment: The Ecological, Atmospheric, and Climatic Importance of Vegetation Fires. John Wiley & Sons, New York. Cleaves, D. A., and T. K. Haines. 1995. Regulation and liability risk: influences on the practice and the pricetag of prescribed burning. Pages 165-183. Environmental regulation and prescribed fire: legal and social challenges. Florida State University, Tampa Airport Hilton at Metrocenter, Tampa, Florida. Cleaves, D. A., J. Martinez, and T. K. Haines. 2000. Influences on prescribed burning activity and costs in the National Forest system. General Technical Report SRS-37. U.S. Department of Agriculture, Forest Service, Southern Research Station, Asheville, NC. Clewell, A. F. 1989. Natural history of wiregrass (Aristida stricta Michx., Gramineae). Natural Areas Journal 9: 223-233.

149 Cordell, H. K., J. C. Bliss, C. Y. Johnson, and M. Fly. 1998. Voices from southern forests. Pages 332-347 in K. G. Wadsworth, ed. The Sixty-third North American Wildlife and Natural Resources Conference. Wildlife Management Institute, Orlando, Florida. Costa, R. 1995a. Red-cockaded woodpecker recovery and private lands: a conservation strategy responsive to the issues. Pages 67-74 in D. L. Kulhavy, R. G. Hooper, and R. Costa, eds. Red-cockaded Woodpecker: Recovery, Ecology and Management. Center for Applied Studies in Foresty, College of Forestry Stephen F. Austin State University, Nacogdoches, Texas. Costa, R. 1995b. Red-cockaded woodpecker species recovery: agencies' and individuals' challenges, responsibilities and opportunities. Pages 22-27 in D. L. Kulhavy, R. G. Hooper, and R. Costa, eds. Red- cockaded Woodpecker: Recovery, Ecology and Management. Center for Applied Studies in Forestry, College of Forestry Stephen F. Austin State University, Nacogdoches, Texas. Crist, P. L., T. W. Kohley, and J. Oakleaf. 2000. Assessing land-use impacts on biodiversity using an expert systems tool. Landscape Ecology 15: 47-62. Dillman, D. A. 2000. Mail and Internet Surveys: The Tailored Design Method Second Edition. John Wiley & Sons, Inc., New York. Estill, E. 1999. Croatan National Forest: land and resource management plan: [Proposed]. Management Bulletin R8; 85A. U.S. Department of Agriculture, Forest Service, Southern Region, Atlanta, GA. Fried, J. S., G. J. Winter, and J. K. Gilless. 1999. Assessing the benefits of reducing fire risk at the wildland- urban interface: a contingent valuation approach. International Journal of Wildland Fire 9: 9-20. Frost, C. C. 1993. Four centuries of changing landscape pattern in the Longleaf Pine Ecosystem. Pages 17-43 in S. M. Hermann, ed. The Longleaf Pine Ecosystem: Ecology, Restoration and Management. Tall Timbers Research, Inc., Tallahassee, Florida. Gilliam, F. S., and W. J. Platt. 1999. Effects of long-term fire exclusion on tree species composition and stand structure in an old-growth Pinus palustris (Longleaf pine) forest. Plant Ecology 140: 15-26. Glitzenstein, J. S., William J. Platt, and Donna R. Streng. 1995. Effects of fire regime and habitat on tree dynamics in north Florida longleaf pine savannas. Ecological Monographs 65: 441-476. Haines, T. K., and D. A. Cleaves. 1999. The legal environment for forestry prescribed burning in the south: regulatory programs and voluntary guidelines. Southern Journal of Applied Forestry 23: 170-175. Hauenstein, E. B., and W. C. Siegel. 1980. Air quality laws in southern states: effects of prescribed burning. Southern Journal of Applied Forestry . Johnson, R., and D. Gjerstad. 1998. Landscape-scale restoration of the longleaf pine ecosystem. Restoration & Management Notes 16: 41-45. Outcalt, K. W., and R. M. Sheffield. 1996. The longleaf pine forest: trends and current conditions. Resource Bulletin SRS-009. U.S. Department of Agriculture, Forest Service, Southern Research Station, Asheville, NC. Platt, W. J., G. W. Evans, and S. L. Rathburn. 1988. The population dynamics of a long-lived conifer (Pinus palustris). The American Naturalist 131: 491-525. Theobald, D. M., N. T. Hobbs, T. Beary, J. A. Zack, T. Shenk, and W. E. Riebsame. 2000. Incorporating biological information in local land-use decision making: designing a system for conservation planning. Landscape Ecology 15: 35-45. Wade, D., and K. Outcalt. 1999. Prescription fire to manage southern pine plantations--damned if you do, damned if you don't. TAPPI International Environmental Conference, Nashville, TN. Ware, S., C. Frost, and P. D. Doerr. 1993. Southern mixed hardwood forest: the former longleaf pine forest. Pages 447-493 in W. H. Martin, S. G. Boyce, and A. C. Echternacht, eds. Biodiversity of the Southeastern United States: Lowland Terrestrial Communities. John Wiley & Sons, Inc., New York.

150 THEY'RE BA--ACK -- HARDWOOD SPROUTING 3 YEARS AFTER RESTORATION HARVESTING IN LOWER ALABAMA

John S. Kush (Auburn University School of Forestry & Wildlife Science, 108 M. White Smith Hall, Auburn University, AL 36849) J. Morgan Varner, III (Interdisciplinary Ecology Program, Box 118526, University of Florida, Gainesville, FL 32611-8526)

ABSTRACT: The Flomaton Natural Area is an old-growth longleaf pine stand located in south Alabama. The stand had experienced 40+ years of fire suppression prior to restoration efforts beginning in 1995. Several studies are being conducted as part of the restoration efforts. One of these is a study following the fate of hardwood stems. The first two prescribed fires implemented in the stand were effective on stems less than 3-inches in diameter. However, they had no effect on hardwood stems greater than 4-inches in diameter. After a fuelwood operation in 1997 removed most of the hardwoods, sprouting of stumps was examined. Sprouting was greater within the more shade tolerant species, southern magnolia, and American holly than it was within the oaks. In addition, the smaller the stump was, the more likely it is going to re-sprout. If the use of fire cannot happen carefully and often enough then other options must be considered.

INTRODUCTION Longleaf pine ecosystems evolved under a chronic fire regime. The presence of fire is required to keep them healthy. When fire is removed from the system hardwood species quickly invade, increasing stand density, shading out herbs, grasses, and pine regeneration, increasing competition with pines, and impeding fire movement. To remedy this process, restoration activities can be instituted to restore ecosystem functioning and processes.

How to go about the restoration process is a question not easily answered. With the declining longleaf pine acreage, great care must be undertaken in trying to achieve restoration objectives. It may be necessary to take several actions: various types of thinnings, removing a hardwood mid- and understory, increasing fine fuels, use of herbicides, re-introducing fire, and many others.

In fire-suppressed stands, harvesting the hardwood midstory seems a necessary action. The outcome of sprouting hardwoods, however, is an ominous unknown. To examine the issue of sprouting, the fate of hardwoods during the restoration process at the Flomaton Natural Area (FNA) has been followed. Sprouting of hardwood stems was examined after two prescribed fires had been conducted and a fuelwood operation, which cut nearly all hardwood, stems.

METHODS Study Site The Flomaton Natural Area (FNA) is a 60-acre virgin, old-growth longleaf pine stand located in the city limits of Flomaton in western Escambia County, AL. By 1995, the stand had experienced over 40 years of fire suppression. The Alger-Sullivan Lumber Company, one-time owner, preserved this stand through the first half of the century. As part of the preservation effort, the stand was regularly control burned until about 1950, after which the stand had remained unburned. The stand is now owned by International Paper (formerly Champion International Corporation).

In 1993, 32 1/5-acre research plots were established within the stand. Data recorded for every tree on the plots included: species, azimuth and distance from plot center, diameter at breast height (DBH), crown and total height, and litter depth at the base of the stem.

RESULTS Pre-restoration data A total of 24 tree species were recorded. Longleaf pine was the dominant species based on the density and the basal area of the overstory trees. However, the absence of fire allowed oaks, specifically water oak, laurel oak, and southern red oak to become a major component of the stand. These three oaks accounted for 37% of tree density and 15% of the basal area for species in the overstory. In this midstory, oaks accounted

151 for 45% of the stems compared to less than 30% for longleaf pine and would potentially replace longleaf pine in the overstory as large tree mortality occurred.

The oaks dominated the sapling-size stems in the FNA. They accounted for 56% of the density and basal area of this size class. Longleaf pine had dropped to 5% of the stems and 11% of the basal area. Shade tolerant species like American holly, flowering dogwood, and southern magnolia had increased in importance. The other pines, which were part of the overstory, had very few stems in the understory. Unless restoration efforts were undertaken, the FNA would have eventually converted into a mixed hardwood stand.

The restoration process in the Flomaton Natural Area • Fire was re-introduced to approximately half the stand in January, 1995 and the other half in April. • The same burning regime was followed in 1996. • In April-May, 1996, a fuelwood operation was conducted by the Easterling Brothers of Brewton, AL. They removed 1350 tons of hardwood chips and inflicted very little damage to the residual stand. • In June 1997, a spring burn was used in an effort to reduce and eventually eliminate hardwood sprouting. • Plans were to implement spring burns in 1998, 1999, and 2000 but dry weather conditions made this too dangerous. • Plans are to burn the stand during the 2001 winter.

Hardwood sprouting Stumps of stems that were 4 inches in diameter and larger in 1996 were re-surveyed in 1999, or three years following the fuelwood operation. Across all species and size classes, 43.9% of stumps sprouted. The more shade tolerant species- American holly and Southern magnolia- had a significantly higher percentage of sprouting than did the other species. The only hardwood species considered “typical” fire-maintained longleaf pine stands – Southern red oak – had the lowest success sprouting. This fact, and the lower than average levels of sprouting observed in flowering dogwood and any other mid-story species leads us to formulate a fire-exclusion hypothesis that is applicable to restoration planning.

GUIDELINES FOR HARDWOOD REMOVAL IN LONGLEAF PINE RESTORATION PROJECTS 1. Evaluate Hardwoods by Species - EARLY! Different species respond differently to different treatments. In response to harvesting in the FNA, Southern red oak sprouting was very poor (24%) vs. other species that sprouted approximately 80%. Also, other important species-specific characteristics may vary (susceptibility to season of fire, litter moisture, canopy leaf area, potential toxicity to other plants, etc…). Also, if masts are important in your objectives, retaining heavy mast producing species may be beneficial. Watch Non-Native species (Chinese tallow, Chinaberry, and a cast of others)- they can quickly invade and alter the disturbed sites that are often created during restoration activities. Overlooking or downplaying their presence will be expensive and time consuming. 2. Evaluate Hardwoods by Size - As we found at the FNA, sprouting was dependent on tree size. Smaller trees sprouted best, with a strong loss in sprouting vigor above ~ 6-inches DBH. However, these large trees provide a heavy seed input, so may represent the biggest danger. Larger trees may also be marketable, providing much-needed income to dollar scarce restoration projects. 3. Consider a Diverse Array of Hardwood Control Methods - At the FNA, we have employed a variety of measures, each meeting site-specific goals. Specific methods we employed were: harvesting, burning, herbiciding (in adjacent problem areas), bush-hogging, and hand-girdling and felling. These treatments, and many others, used in combinations can be used effectively to control problem species. 4. Don’t Forget about the Pine Overstory- Any restoration effort must emphasize the retention of the pine overstory and the valuable needle litter it provides as fuel. Pine needle litter carries fire to kill hardwoods, encourage grassy and herbaceous plant establishment, and expose soil for longleaf pine seedling germination. If hardwoods are to be harvested, operators must be supervised carefully to avoid damage (basal wounds, soil compaction, etc…) to the residual stand. In the absence of an overstory, hardwood sprouting and spread of invasive species is facilitated.

152 ACKNOWLEDGEMENTS The authors wish to thank USDA Forest Service - Southern Research Station, Champion International Corporation (International Paper), Alabama Forestry Commission, The Nature Conservancy, and the Alabama Natural Heritage Trust of the Alabama Department of Conservation and Natural Resources. In particular, we would like to thank Ron Tucker, George Ward, Bill Thompson, Richard Sampson, Chadwick Avery, and Bobby Chamlee for their work over the past year. The major objectives for the Flomaton Natural Area are to restore, monitor, and manage the stand as an old-growth longleaf pine habitat and to conduct biological research that will provide information on old-growth dynamics and techniques for future ecological restoration work.

153 RESTORATION AND MANAGEMENT OF THE ROY E. LARSEN SANDYLAND SANCTUARY, HARDIN COUNTY, SILSBEE, TEXAS

Wendy J. Ledbetter (The Nature Conservancy, P.O. Box 909, Silsbee, Texas 77656)

ABSTRACT: The Nature Conservancy’s Roy E. Larsen Sandyland Sanctuary serves as a key ecological site for protection of the biodiversity of Southeast Texas and longleaf pine conservation efforts. This 5,561-acre project located in Silsbee, Texas lies in the heart of the Big Thicket 20 miles north of Beaumont. The preserve is known for its assemblage of plant communities and species; harboring representative examples of beech-magnolia forests, wetland pine savanna, hardwood bottomland forests, mixed pine hardwood forests, bald cypress-water tupelo swamps and longleaf pine uplands. Ephemeral streams, and ponds, backwater sloughs and acidic baygalls intermingle among these communities.

In 1994, a 3,200-acre buffer area surrounding the core preserve was created by a donation from Temple- Inland Forestry, Inc. of land and a conservation easement. These lands are cooperatively managed with a program of compatible forestry that maintains important ecosystem patterns and processes, and ensures the long-term viability of plant and animal populations. This buffer contains areas that were once historically longleaf pine flatwoods, one of the rarest and most biologically diverse communities in the region.

Slash pine (Pinus elliottii), a non-native species to Texas is being selectively harvested. Natural regeneration and reforestation of longleaf pine is being conducted using West Gulf Coastal Plain seed sources. Use of growing season fires and herbicide has been effective in hardwood species control. Restoration of groundcover species is of particular signficance with over 727 species occurring on the project. Included in conservation efforts is the largest protected population of Texas trailing phlox (Phlox nivalis ssp. texensis).

PROJECT DESCRIPTION The Nature Conservancy’s Roy E. Larsen Sandyland Sanctuary serves as a key ecological site for protection of the biodiversity of Southeast Texas and longleaf pine conservation efforts. This 5,561-acre project located in Hardin County. The preserve is located in the Big Thicket region of the West Gulf Coastal Plain, 20 miles north of Beaumont, Texas.

The preserve is known for its assemblage of plant communities and species; harboring representative examples of beech-magnolia forests, wetland pine savanna, hardwood bottomland forests, mixed pine hardwood forests, bald cypress-water tupelo swamps and longleaf pine woodlands. Ephemeral streams, and ponds, backwater sloughs and acidic baygalls intermingle among these communities.

As in virtually all of the southeastern United States virgin longleaf pine (Pinus palustris) was harvested by the 1930’s. Second growth stands of longleaf still exist today, but many were replaced by the introduction of slash pine (Pinus elliottii), a non-native species to Texas, and loblolly pine (Pinus taeda). In some areas slash pine has stagnated due to poor sandy soils and lack of nutrients. Longleaf, better adapted for these extreme conditions, has the potential to grow more competitively.

Additionally these monoculture stands are exhibiting fusiform rust and are subject to insect infestation. Shading from the dense cover of row plantations will not provide the ideal conditions for many of the region’s specialized plants that prefer the open conditions of the forest floor once provided by naturally occurring stands of longleaf pine.

A restoration and management program for restoration of the longleaf pine ecosystem is presented. This comprehensive long-term program will involve the application of fire, harvesting, reforestation, hardwood control, and forest pest management. Because of the diversity of the site-specific treatments for individual stands and burn management units is required.

FIRE MANAGEMENT The application of fire is the most important management activity used at the preserve for longleaf pine restoration. Prescribed fire is used to mimic the impact that historically lightning fires had on the landscape.

154 Fire is used to reduce fuel loads, control/eliminate hardwood and shrub encroachment, thin dense stands of pine, cause mortality of non-longleaf seedlings, and expose mineral soil for longleaf pine regeneration. A variety of herbaceous groundcover species also appear to be benefit by the use of fire with the project’s 727 plant species.

Fire suppression policies of the twentieth century created dense fuel loads and unnatural shrub thickets. Although there are other alternatives to the use of fire, it is the effects of the physical process of fire that is of significance. An uneven-aged forest with widely spaced-trees of multiple age classes, longleaf regeneration small groupings of hardwood trees, a dominant ground cover of little bluestem (Schizarium scoparium) and a diversity of herbaceous species is considered as a fair representation of successful restoration.

A fire management program was established in 1977. A prescribed burning regime has been established for 1500 acres of the preserve. Prescribed burns are conducted by a statewide burn crew comprised of Conservancy staff and volunteers. Additional support has been provided by the United States Fish & Wildlife Service -Anahuac National Wildlife Refuge, McFaddin National Wildlife Refuge, Big Thicket National Preserve, Texas Parks and Wildlife and the Texas Forest Service.

Growing season burns are utilized to have the greatest impact on hardwood shrub encroachment. Critical periods of the year for the production of longleaf pine are considered. Trees are monitored in the spring months for apical bud growth and in the summer months for insect infestation and stress from drought conditions.

Burning can assist in initial fuel reduction and is scheduled in the late fall to early spring while fire behavior correlates with cooler, wetter conditions. Once fuels are reduced to a manageable level, burns are conducted in the spring to summer months to have a greater impact on hardwood control.

In most cases prescribed burns can be conducted using natural firebreaks such as waterways, ponds, wetland vegetation, or saturated soil conditions. In addition anthropogenic features breaks such as railways, and mowed right-of-ways can be utilized for firebreak preparation.

HARVESTING In some cases the use of fire alone is not sufficient to eliminate the existing slash pine stands. In addition, hastening the removal of these stands and replacing them with longleaf pine is a project restoration goal. Therefore, commercial harvesting and cut- and -leave operations are being used to accelerate the restoration process.

A timber stand inventory and analysis was conducted in 1997-1998 by a forestry consultant using Conservancy monies as well as funds awarded from the Rodney Johnson -Katherine Ordway Stewardship Endowment. This analysis is beneficial in that it provides a detailed picture of the composition and structure of individual stands. From this information harvest and planting recommendations can be formulated.

Formerly owned by the timber industry, the core tract of the Roy E Larsen Sandyland Sanctuary currently contains remnant slash pine and loblolly pine plantations in areas that historically consisted of longleaf pine. These plantations are being commercially harvested where timber operations can be conducted without excessive damage to the site’s components and where harvesting is deemed cost-effective.

Past commercial operations on the preserve have involved the use of a shear, loader and large-tired skidders. Since some harvest operations are of less acreage than large-scale commercial plantations, securing an operator whom is willing to work on a smaller scale is important.

The Conservancy first evaluates the sensitive areas of the potential operation and identifies location of any elements that need protection. Conservancy staff mark unit boundaries, sensitive areas and trees that are to remain. Once the area is subjected to harvest, additional inventories are performed to detect any additional occurrences of targeted species.

155 The target for initial thinning operations is approximately 30 to 60 basal area of pine. The residual stand secures pre-harvest longleaf and shortleaf. Additional loblolly and slash pine may be left to meet the target basal area. These residual trees continue to provide adequate fuel for additional burning, shade and moisture for groundcover species and habitat for wildlife. In some stands large loblolly pines (---cm dbh or greater) are being kept to be used as potential insert trees for red-cockaded woodpeckers; given the possibility of future participation in the U.S. Fish & Wildlife Service’s Safe Harbor Program.

CUT AND LEAVE OPERATIONS Cut and leave operations are utilized in areas where there are sensitive soil and water conditions, inaccessible road conditions or in areas where commercial operations would not be cost- effective.

REFORESTATION In areas that have been identified as having a lack of longleaf seedtrees, commercial seedlings are being used to supplement natural regeneration. Seed is secured from East Texas and western Louisiana sources to ensure genetic stock is derived from the West Gulf Coastal Plain. Stands where natural regeneration is not occurring are targeted for supplemental planting. Seedling survival has varied and been subject to drought conditions during the past two years. Funding through the Global ReLeaf Program has provided additional funds to secure containerized seedlings, which are more expensive than bareroot seedlings. Containerized seedlings will be available for planting in the Fall and Winter of 2000 using local contract crews recommended by the Texas Forest Service and Temple-Inland. It is anticipated that the survival rate of these seedlings will be greater than 80% .

RARE SPECIES MANAGEMENT As with restoration efforts in many ecosystems, the longleaf pine is not the only species benefiting from management activities. The Sandyland project is most noted for the ongoing success of enhancing the population of the Texas trailing phlox (Phlox nivalis ssp. texensis). This federally endangered species, occurs in the sandy soils of the longleaf pine woodlands. Initial surveys of Sandylands located 40 of these plants. With years of prescribed burning and additional surveys 400 plants have been identified in two distinct populations. The preserve harbors the largest protected population of this species and a critical site for its recovery. Other rare plants benefit from the application of fire include scarlet catchfly (Silene subciliata) and gaillardia (Gaillardia aestivalis var. winkleri).

HERBICIDE Enhancement of herbaceous groundcover species and control of hardwood encroachment are both restoration goals of the Sandyland project. Fire suppression has allowed hardwood shrubs to increase, altering the microclimate conditions for herbaceous growth. Control of hardwood growth is needed, but broadcast application of herbicide could potentially harm the wide diversity of groundcover species. The specific effects of herbicide is unknown for many species, particularly those that are considered to be in limited distribution or to be of ecological significance.

The amount of hardwood to remain and its configuration is debated, but for the purpose of this project hardwood coverage less than 20 BA is the identified target. Some larger individuals (8 inches DBH or more) of post oak (Quercus stellata) and other hardwood species are left in stands. To control hardwood shrub encroachment herbicide is being used as a basal bark treatment predominately on bluejack oak Quercus incana, and American Holly (Ilex opaca), and Yaupon Holly (Ilex vomitoria) in upland areas. Application of a 20% mixture of Garlon 4 and JLB Oil with backpack sprayers has proven highly effective with complete mortality. Since this management treatment is labor intensive, application to smaller trees (< 4 inches DBH) is considered only where hardwoods are extremely dense and fire has proven ineffective in impacting the hardwood population. Using herbicide in combination with prescribed fire appears to have an even greater impact than using either treatment alone.

SOUTHERN PINE BEETLE CONTROL Southern pine beetles and other insect infestations are monitored on the ground and by Texas Forest Service reconnaissance flights. In the event of an outbreak the Conservancy monitors the spot and determines a course of action. Salvage operations and cut-and –leave operations are two alternatives used when infestation occurs. In some instances residual longleaf pine may be threatened when infestations occur in dense slash or

156 loblolly pine plantations. When an outbreak does not threaten commercial stands on the Temple-Inland easement, privately-owned lands, or native natural stands the infestation is allowed to run its course, assisting in the decline of the slash pine on site.

PARTNERSHIP FOR ECOSYSTEM CONSERVATION Temple-Eastex, Inc established the Roy E. Larsen Sandyland Sanctuary in 1977 by a donation of 2,138 acres. This gift has now become a core of a larger cooperative project whose goals include conserving the longleaf pine ecosystem and developing a compatible forestry program.

In 1994 The Nature Conservancy and Temple-Inland Forestry Products Corporation, Inc. renewed their commitment to conservation by placing 2,805 acres surrounding the preserve under a conservation easement. Both parties agreed to protection of sensitive elements, no commercial or residential development, preservation of bottomland hardwoods and to develop a compatible longleaf pine forestry where site conditions warranted. Income derived from hunting leases is an additional value to Temple-Inland.

A five-year management plan identifies treatments for individual stands of approximately 1,800 acres of slash and loblolly plantations. Temple-Inland staff and Conservancy staff work together to identify rare species, sensitive habitat, delineate streamside management zones, design entry points and access to the tract, consult with contract loggers, identify log set sites, and determine equipment and harvest methods.

The Sandyland project is an innovative agreement between a fiber products company and a private conservation organization. The success of conserving the longleaf pine ecosystem is dependent on the willingness of both private and public and industrial and non-industrial partners to work cooperatively. It is hoped that this project will serve as a demonstration site for the creation of new projects and partnerships will be established for the conservation of the longleaf pine ecosystem

157 DON’T MISS THE FORBS FOR THE TREES: PRELIMINARY RESEARCH PROPOSALS FOR THE STUDY OF LONGLEAF PINE SAVANNA’S IN EASTERN LOUISIANA

Van Lopez (University of Southeast Louisiana, Hammond, LA) Arie Roth (University of Southeast Louisiana, Hammond, LA)

ABSTRACT: The westernmost edge of the Longleaf Pine ecosystem of the East Gulf Plain lies in the Florida parishes of eastern Louisiana. The Nature Conservancy is working to conserve the last remnants of this ecosystem in Louisiana at their Lake Ramsay, Abita Springs, and Talisheek preserves. The research team of the Edward G. Schleider Endowed Chair in Environmental Studies at Southeastern Louisiana University propose several to support TNC’s efforts in this ecosystem.

We will collect preliminary species/area data from the Lake Ramsay Preserve, present sites with three different ages since last burn. We propose three studies of these sites. The first is a study of competition in partially restored longleaf pine savannas. We will remove dominant species from plots, and measure the species richness response to this perturbation.

The second project will be a descriptive classification and ordination of the Talisheek site. We will measure four subjectively selected habitats to answer questions about the relationship of species composition to fertility, moisture, and other environmental factors. We will also address questions about which habitat has the most species and most rare species.

Finally, we will study the stability of the ecosystem at Talisheek. We will disturb the system with fire, and test this habitat’s level of resistance (e.g.,stability) and relate this to richness. We will also test how resistance varies with each of the environmental factors measured in project two.

158 RESTORATION OF WET LONGLEAF PINE SAVANNA IN SOUTHEAST LOUISIANA

Richard P. Martin (The Nature Conservancy, P. O. Box 4125, Baton Rouge, LA 70821) Susan Carr (The Nature Conservancy, P. O. Box 4125, Baton Rouge, LA 70821)

ABSTRACT: The Abita Creek Flatwoods Preserve is located in the East Gulf Coastal Plain ecoregion of Southeast Louisiana. The Louisiana Field Office of The Nature Conservancy acquired this 794-acre tract in 1996, as an addition to the Southeast Louisiana Pine Wetland Mitigation Bank. Our initial stewardship program at Abita preserve has focused on restoration of species-rich longleaf pine savanna via commercial logging of slash pine, growing season prescribed fire, and planting longleaf pine seedlings. In addition to affecting rapid removal of invasive slash pine, the timber sale generated income to partially defray acquisition costs. Our primary research interests are the initial effects of logging, and the subsequent recovery rate of natural community composition and structure. Specific monitoring questions are: 1) What are the effects of commercial logging on species richness, abundance, and composition and 2) how does logging coupled with fire treatments compare to “fire only” treatments. Here, we present some preliminary results that address the first question.

159 EFFECTS OF STOCK TYPE, FALL NURSERY FERTILIZATION AND ECTOMYCORRHIZAL INOCULATION ON SURVIVAL OF LONGLEAF PINE (PINUS PALUSTRIS MILL.) SEEDLINGS PLANTED ON LIGNITE MINESPOIL

Mary Anne McGuire (Warnell School of Forest Resources, University of Georgia, Athens, GA 30602) Emily Goodwin (Arthur Temple College of Forestry, Stephen F. Austin State University, Nacogdoches, TX 75962) Hans Williams (Arthur Temple College of Forestry, P.O. Box 6109, Stephen F. Austin State University, Nacogdoches, TX 75962)

ABSTRACT: Bareroot and container longleaf pine (Pinus palustris Mill.) seedlings were planted on lignite minespoil in east Texas in 1996 and 1997. Effects of stock type, fall nursery fertilization (nitrogen and phosphorus) and ectomycorrhizal inoculation on seedling morphology, mineral nutrition, physiology and survival were assessed. Bareroot seedlings had larger root collar diameters and greater root weight and shoot weight than container seedlings. Container seedlings had higher root to shoot ratios. Fertilizer treatments had no statistically significant effect on seedling morphology. Bareroot seedlings had greater foliar nitrogen concentrations at time of planting than container seedlings. Seedlings fertilized in the fall with nitrogen had greater foliar N concentrations at time of planting. Root growth potential was significantly greater for container seedlings. Needle water potential, stomatal conductance and transpiration tended to be greater for container seedlings during the first growing season following planting. Fertilizer treatments had no statistically significant effect on root growth potential and water relations. First year survival was greater for container seedlings following the 1996 and 1997 plantings. It appears that successful regeneration of longleaf pine on lignite minespoil sites in east Texas can be accomplished more reliably with container seedlings than with bareroot seedlings.

160 SUSTAINABILITY AND PRODUCTIVITY OF SOUTHERN PINE ECOSYSTEMS: A THEMATIC FRAMEWORK FOR INTEGRATING RESEARCH AND BUILDING PARTNERSHIPS

Charles K. McMahon (USDA Forest Service, Southern Research Station, 520 Devall Dr, Auburn, AL) James P. Barnett (USDA Forest Service, Southern Research Station, 2500 Shreveport Highway, Pineville, LA)

ABSTRACT: In 1997, the USDA Forest Service Southern Research Station (SRS) published a Strategic Plan that formed a framework for addressing the Sustainability of Southern Forest Ecosystems. Six crosscutting themes were identified to facilitate research integration and partnership building among the widely dispersed SRS research work units. The Sustainability and Productivity of Southern Pine Ecosystems theme outlined in this paper, allows us to identify critical issues, information gaps, and research needs for ecologically sound, economically viable, and socially acceptable management of the southern pine and pine- hardwood ecosystems.

Introduction: In 1997, the USDA Forest Service Southern Research Station (SRS) published a Strategic Plan that formed a framework for addressing the Sustainability of Southern Forest Ecosystems (USDA Forest Service 1997). With our partners and other resource users we identified issues and needs that will determine the direction and effectiveness of SRS research and development programs. The SRS strategic plan forms a basis for a collaborative multi-disciplinary approach to these programs. The plan describes three broad research goals:

1. Measuring and monitoring forest resources. (What do we have?) 2. Understanding ecosystem structure, function, and processes. (How does it work?) 3. Ensuring environmental quality and sustainable productivity. (How can we use it without losing it?)

In order to achieve these goals, six operational crosscutting themes (CCT’s) were established to integrate the work of our highly decentralized organization, build partnerships, and develop products that meet customers’ needs. Four crosscutting themes focus on predominant manifestations of the South’s forest resources, while two themes focus on the sustainable management of all forest ecosystems in the Southern Region. The themes include:

1. Southern Appalachian Ecosystem Research and Sustainability 2. Sustainability and Productivity of the Interior Highlands Ecosystem 3. Sustainability and Productivity of Southern Pine Ecosystems. 4. Ecology and Management of Forested Wetlands, Bottomland Hardwoods, and Riparian Zones 5. Landscape and Regional Integrated Assessment and Modeling 6. Inventory and Monitoring

These themes represent a framework for organizing research and development activities in a way that lowers traditional disciplinary and institutional boundaries. They are not highly structured formal programs. They are dynamic and flexible; enabling the SRS research community to adapt to evolving customer needs and respond to emerging issues. Over 20 SRS research work units are contributing scientific support and financial resources to the challenges embedded in the themes. In addition, each research unit continues to support traditional SRS research and development studies and programs not covered by the themes.

OBJECTIVES The Sustainability and Productivity of Southern Pine Ecosystems theme provides the basis to identify critical issues, information gaps, and research needs for ecologically sound, economically viable, and socially acceptable management of the southern pine and pine-hardwood ecosystems. Equally important, the theme development process provides a mechanism to bring together scientists, managers, and stakeholders, who can reach consensus on priorities. This theme is very broad, both technically and spatially. Key elements apply too much of the SRS research program, as well as to the programs of forest industry, southern universities, and some private research institutions. This paper is an abridged version of the pine theme booklet (USDA

161 Forest Service 2000) prepared by the members of the theme steering committee.1 The booklet is now available on-line from the SRS web site http://www.srs.fs.fed.us

DEFINITION In 1987, the World Commission on Environment and Development defined sustainability as “meeting the needs of the present without compromising the ability of future generations to meet their own needs.” Southern pine ecosystems include those pine and pine-hardwood forests that are located within the southern coastal plain and Piedmont areas. Sustainable productivity in the context of forest ecosystems relates to the full spectrum of resources that people value in forests including wood fiber, recreation use, water yield and quality, abundance and diversity of flora and fauna, and other valuable resources. Managing these resources for their sustainability will require us to recognize human actions. In the context of ecosystem sustainability, forests should produce desired resource values, user products, and services in ways that maintain ecosystem health (Burkett and others 1996).

Values and Demands for Resources: The South supplies 67 percent of the Nation’s pulpwood, 50 percent of its plywood, 40 percent of its hardwood lumber, and 33 percent of its softwood lumber. Timber is the region’s highest valued crop, representing an annual economic value of $90 billion. In an average year, removal of wood products in the South totals 8.9 billion cubic feet, only about 4 percent of which comes from national forests. In addition to being an essential source of wood products, the region’s forested lands support a robust recreation business; they provide clean air; supply abundant water for domestic, agricultural, and industrial uses, as well as recreation; maintain diverse habitats for plants and animals; and serve as a potential sink for atmospheric carbon. Our human population growth has been accompanied by increased demands for forest resources and a chorus of opinions about how America’s forests should be managed. Expanding urban populations have clear expectations of environmental quality and the availability of resources. Coupled with the growing affluence in American society is a growing interest in conserving and enhancing soil, air, water, wildlife, fish, and recreation resources. The South can expect a greater demand for pulpwood, lumber, and other products, as well as outdoor recreation opportunities. Most pressures on forest resources will be felt on private ownership, which constitutes nearly 90 percent of the South’s forests

Barriers to Sustainability: Trends in soil and fertility losses, epidemic levels of insect pests and pathogens, losses of threatened, endangered, or sensitive (TES) species, encroachment of exotic weeds, forest fragmentation, and other problems affecting forest health—all complicate the challenge of ensuring southern pine ecosystem sustainability. The history of most forest management practices is short; there has been little documentation of their long-term effects on forest sustainability. For example, loss of soil productivity can result from the repeated removal of biomass from the forest floor. Annual losses may be small, but cumulative losses may have significant impacts. We need knowledge of the long-term effects of forest management practices on basic forest resources, such as soil productivity, water quality and quantity, biodiversity, and wood production. A more obvious risk to sustainability occurs when forest ecosystems are lost to encroachment from urban development, highway and power line expansions, and other human activities. As the amount of forest land decreases, societal demands exert even greater pressure on the forests that remain. Other barriers to sustainability result from dramatic environmental changes that are brought about partly by rapid population growth and urbanization, and partly by the domestic and international demands for resources. Human demands, once primarily for wood products, now include non-traditional products, and recreational pursuits.

Management Challenges: In the South, as elsewhere, there are competing demands for limited resources. Most notably, intensive management for timber and other forest resources is sometimes in direct conflict with TES habitat protection. Allocation of the resources that sustain us has become a critical issue, especially in the South, where 90 percent of the forests are within private ownership. Most small forest landowners in the 13 southern States have little capacity to conduct research to improve resource sustainability.

1 SRS Sustainability and Productivity of Southern Pine Ecosystems theme steering committee: James Barnett, Rod Busby, Marianne Burke, Floyd Bridgewater, Don English, James Hanula, Nancy Herbert, Kurt Johnsen, Brad Kard, Kier Klepzig, Charles McMahon, Jeff Prestemon, Tim Rials, Bob Rummer, Mike Shelton, Ron Thill, Tom Waldrop, Joan Walker, and Nancy Walters. Theme co-leaders are Barnett and McMahon.

162 Nonetheless, by applying the knowledge and technology developed by government, industry, and university research, they can benefit substantially. In addition, managers of privately owned industrial forest lands could apply the results of cooperative research to sustain wood production and other forest benefits. In response to booming populations and economies, the global demand for affordable construction materials, paper products, fuelwood, and wood chemicals is growing exponentially. Recent timber-harvest reductions in the American West have brought unprecedented pressures on the South, especially its private forest lands.

The Montreal Process (MP) is a process that describes a comprehensive set of seven criteria for forest conservation and sustainable management that is being used by the international forestry community. The MP evolved from the Working Group on Criteria and Indicators for the Conservation and Sustainable Management of Temperate and Boreal Forests, which was convened in Geneva, Switzerland, in June 1994 (Canadian Forest Service 1995). The MP criteria were designed to provide: (1) a common understanding of what is meant by sustainable forest management, and (2) a common framework for describing, assessing, and evaluating a country’s progress toward sustainability at the national level. Although the criteria are not intended to assess sustainability at a regional or forest level, they do provide an internationally recognized template for categorizing or grouping research questions and needs. We have used the seven criteria to sort the broad issues and research questions about southern pine ecosystems.

MP Criterion #1: Conservation of Biological Diversity The Southern United States is nearly unmatched in biological and genetic diversity. On this rich landscape a long history of intensive and extensive human use has occurred. Our growing human population, coupled with the varied land-use practices has contributed to; over-exploitation of plant and animal resources, simplification of ecosystems, alteration of natural processes, (e.g., reducing the frequency and extent of wildfires), habitat fragmentation, changes in genetic variation and dynamics, and habitat degradation and loss. The combination of wide biodiversity at multiple scales and the extensive uses made of southern landscapes complicate most efforts to conserve and restore diversity. There are few relatively undisturbed ‘reference’ sites within southern pine ecosystems, making it hard to know which elements have been lost or to interpret the patterns that remain. The status and distribution of rare species or rare genotypes are particularly difficult to determine. Some historical uses have brought changes, e.g., depleted soils that are not typical of the component species’ recent evolutionary history. Also, fire-exclusion policies have removed a controlling element from many pine ecosystems, resulting in changed floral, faunal, and structural diversity. Reintroducing the natural disturbance regime probably would not restore diversity. Ecosystem structure and function will first have to be restored. Where complex land ownership patterns and the growing urban/wildland interface are occurring, the use of fire is problematic. The underlying ecological diversity of southern pine ecosystems will make it hard to find general solutions for maintaining and restoring biological diversity. In order to address this criterion for sustainability, scientists from many disciplines must address three basic research goals:

1. Define baseline biodiversity at multiple scales e.g., landscape, stand, within-stand, and within- species (genetic diversity)] across southern pine ecosystems. 2. Determine how alternative forest management practices affect biological diversity at various scales. 3. Determine how biodiversity can be maintained and restored in the southern pine region, and develop economically and environmentally feasible technology to do so.

MP Criterion # 2: Maintaining the Productive Capacity of Southern Pine Ecosystems Southern pine forests have become the world’s largest source of wood fiber; however, we need to consider forest productivity in a much broader context. Recreational opportunities, socio-economic worth, wildlife habitat, and other non-commodity aspects figure prominently in a forest’s value. In addition, some very important questions remain regarding the sustainability of intensive silviculture. Science has shown us how fertilization, control of competing vegetation, and the use of improved genotypes can improve plantation performance; but what are the risks of pushing tree growth to its limit? Will such plantations be vulnerable to damage by pests, or to the effects of severe and infrequent environmental conditions like 100- year droughts? What about the long-term sustainability of soil resources over the course of repeated rotations? Alternative management practices may affect the quality of wood, as well as other forest resources. How does plantation forestry affect wildlife species, recreation and other, less-tangible but very important values? Will such

163 practices indirectly benefit other more-sensitive domestic and international forest ecosystems? How can the productive capacity of the South’s vast non-industrial forestry sector be increased? Are our strategies to increase productivity economically feasible or socially desirable? Maintaining the productivity of southern pine ecosystems will require research that focuses on some key questions:

1. What is the ecosystem’s potential capacity for supplying an array of forest products? 2. What are the limitations to sustained production? 3. How will forest management influence the ecosystem’s long-term productive capacity?

MP Criterion # 3: Maintenance of Forest Ecosystem Health and Vitality As demands for wood and other forest products sharply increase, forestry professionals and non- professionals alike are striving to maintain and restore forest health. They are emphasizing improved forest health because much present-day forest land in the South was farmed long before modern soil conservation practices were known. In many areas, the land was highly eroded or nutrient depleted from heavy use before today’s forests were established. Many areas are succumbing to insect infestation, pathogens, and invasive weed species, any of which may threaten ecosystem health. Problems have come to southern pine forests. For example, nearly 500,000 acres of Florida’s forest land burned in 1998. Although wildfire is a natural disturbance to which ecosystems have adapted, years of fire exclusion have brought unusually high fuel concentrations and, as a result, abnormal fires and fire effects. Other natural disturbance factors, such as hurricanes and southern pine beetles, are affecting the changed forest ecosystems; and those effects may be quite different from what occurs in normal, healthy forest communities. Land managers often have limited knowledge of the influence of disturbance regimes on some cosystem components, as well as their importance to the ecosystem’s overall health and productivity. Other pressures from growing human populations include the fragmentation of forests by road construction, urban sprawl, and changing land use patterns. How, then, do we increase forest productivity while improving forest health? What would landowners, and society as a whole, have to pay to implement such measures, and what benefits could they expect? Fundamental questions about maintaining and enhancing forest health in southern pine ecosystems include:

1. What is the condition of southern pine forests today? Baseline information about forest conditions is needed to ensure that future actions help improve forest health. 2. How do animals (earthworms to deer) and microorganisms affect the health of southern pine ecosystems? The functional role and impacts of only a few organisms in southern forests are understood. 3. How do fragmentation and changing land-use patterns affect southern pine ecosystem function and health? 4. What is the role of major forest disturbances in the overall health and renewal of pine forest ecosystems?

MP Criterion # 4: Conservation and Maintenance of Soil and Water Resources While land managers and property owners try to increase production from a relatively fixed, intensively managed land base, they are often constrained by wetland regulations. Forestry practices in the South are highly manipulative and can affect the soil properties on which sustained productivity depends; soil quality easily can be compromised. Forestry activities may have profound effects on both soil and water, which are closely linked throughout the South. The lean Water Act requires that the impacts of forest practices, on adjacent ecosystems, as well as the managed land itself, be kept to a minimum. Nonetheless, silvicultural operations can influence water quality through sedimentation, hydrologic regimes, changes to channel structure, and biogeochemical processes. If soil and water degradation are to be avoided, we need to better understand the nature of such impacts, as well as appropriate methods for restoring soil and water components of affected ecosystems. Sustaining soil productivity and restoring the productivity of damaged sites reflects a key conservation ethic; and it makes good sense from an economic perspective. As human populations and forest management dramatically increase, soil and water conservation become especially important. Quantifying baseline conditions is a critical first step in developing management practices that will mitigate and improve soil and water conditions, and will help to ensure that forest and aquatic ecosystems provide their bounty for future generations. The fundamental research questions that must be addressed are:

1. What are the baseline conditions of soil and water resources in the southern pine region? 2. What are the interactions among management practices, soil conditions, and water quality?

164 3. What methods can be used to mitigate and improve soil and water conditions in southern pine forests? 4. How does context, i.e., arrangement of different forest types and management regimes on the landscape, affect soil and water resources?

MP Criterion # 5: Maintenance of Forest Contribution to Global Carbon Cycles Some scientists have suggested that forests, and forestry, play an important role in atmospheric carbon sequestration and thereby help mitigate greenhouse gas accumulation and global change. Pine ecosystems in the South, which constitute one of the most important forest assets in the world, potentially could have a significant impact on the world’s carbon cycle. However, the southern pine region is an aggregate of many forest types that are managed at various intensities. Pine plantation silviculture is evolving rapidly, and management tools like fertilization, site preparation, vegetation control, and improved genotypes have greatly influenced net ecosystem productivity (NEP) by reducing rotation ages and by altering soil carbon dynamics. Global warming, by affecting tree growth and function, and by influencing rates of soil respiration, may further influence NEP. In forests managed for wood products, the absolute amount of carbon sequestered from the atmosphere depends on how the harvested wood is used. Wood fiber is processed into any number of products, the life spans of which will vary from months to centuries. Carbon costs, which are associated with management, harvest, transportation, and processing, all contribute to the carbon cycle equation. Given the complexity and scale of such issues, we are not yet able to predict the effects of southern pine forestry on the global carbon budget. Nor can we make policy and management decisions that will allow both competitive industrial forestry and the maintenance of southern pine plantations as an overall carbon sink. The following broad questions about carbon cycling must be addressed.

1. Is the southern pine ecosystem an overall source or a sink of atmospheric carbon? 2. How does forest type influence the ecosystem’s status as a source or sink? 3. What is the influence of different forest management techniques on the long-term status of pine forests as a source or sink? 4. What are potential effects of global climate change on carbon sequestration?

MP Criterion #6: Maintenance and Enhancement of Long-Term Socio-Economic Benefits to Meet the Needs of Societies Southern pine forests provide a diverse set of benefits to landowners, as well as the general public. They are a source of raw materials and income for industrial landowners and wood fiber consumers. They not only provide private, nonindustrial landowners with wood, but also offer American society recreation opportunities, scenic beauty, places that respond to our spiritual needs, and habitat for flora and fauna. Public forestlands also provide these benefits; but, in addition, they provide watershed protection, ecosystem stability, and a stabilizing component in local and regional economies. It is clear, therefore, that the land-use choices of private and public landowners, the demands of forest-product suppliers and consumers, and the many values that we humans place on the natural world will determine the character and extent of the southern pine resource. What is not clear, however, is how we can best interact with the ecosystems upon which we all depend. Southern Research Station scientists and their counterparts in the university community are using applied research to better understand the commodity and non-commodity values derived from private and public forests. Key questions include:

1. How do commodity and non-commodity values affect the amount and character of southern pine resources, and vice versa? 2. How do social and human factors influence management of southern pine forests? 3. What are the potential approaches or strategies that can be used to help limited-resource landowners increase the total value of their forests? 4. What are the relationships between rural communities and the southern pine resource? 5. How best can we assess the non-commodity values of southern pine forests?

MP Criterion # 7: Legal, Institutional and Economic Framework for Forest Conservation and Sustainable Management Management of southern pine forests is tempered by the legal, institutional, and economic framework on which land-use decisions are made. This framework affects the long-term sustainability, productivity, and ecological integrity of the southern pine ecosystem. For example, landowners face a myriad of laws and

165 regulations that directly or indirectly affect their use of southern pine resources. Also, tax, inheritance, and environmental laws can affect the flow of capital into and out of forest management; and these affect investment in long-term forestry. In the South, this framework has implications for the long-term sustainability of ecosystem outputs, timber markets, and other commodity and non-commodity values. Changes in land-tenure patterns are leading to fragmentation and revised management objectives that may, in turn, restrict management options that are or will be available to the landowner. The following questions consider the framework for forest conservation and sustainable management:

1. Policy makers can implement regulations and incentive programs to encourage certain types of forest management; but how effective will these techniques be? How will they affect the resource(s) they are designed to protect? How will the effects differ over time? And what will be their effects on other desired behaviors? 2. What are the effects of recycling, use of non-wood substitutes, national and international competition, and technological change? What values do southern forests produce, and what will be their long-term sustainability? 3. What are the welfare and market implications of sustainable forestry? What will be the long-term, ecological consequences of implementing changes in forestry policies and practices? 4. How are limited-resource southern pine forest owners affected by current or proposed tax, inheritance, or environmental laws, notwithstanding the vagaries of market changes? 5. How do these effects compare with those experienced by other groups? Are laws affecting southern pine forests meeting the policy objectives of lawmakers? How are institutional factors related to long-term sustainability of the southern pine resource?

Examples of Collaborative Research and Development: A number of multi-disciplinary, collaborative research activities are already occurring at the Southern Research Station. Many are designed to encourage additional, cooperative efforts. Here are just four examples that support the Southern Pine Ecosystems theme:

1. Longleaf Pine Restoration and Management In 1994, 17 scientists from 14 SRS research projects joined with employees in the USDA Forest Service Southern Region to develop a broad plan for longleaf pine ecosystem restoration and management. Early team efforts focused on an assessment of current conditions, strategic planning, and partnership building. This team effort provided an opportunity to develop a more integrated approach to longleaf pine ecosystem research. The Station’s long-term, core longleaf research studies were bolstered by Ecosystem Management grant monies; and those funds were used to develop an internal, competitive grant program that would provide seed funding for new studies, or to enhance ongoing studies. By 1997, over 70 manuscripts, abstracts, and posters resulted from this effort. More importantly, a process of collaboration and communication was made available to scientists with common interests and goals. In partnership with the Longleaf Alliance, scientists and managers from across the South now meet regularly to address collaborative research and management strategies related to sustainability of the longleaf overstory and understory communities. The SRS longleaf program addresses important scientific questions related to restoration ecology, fire ecology, smoke management, impacts of silvicultural alternatives on plant and animal communities, improved technology for longleaf overstory/understory regeneration and the socio- economic factors associated with sustainable management of both Federal and non-Federal lands.

2. Long-term Soil Productivity/ Monitoring Productivity and Environmental Quality In 1990, the Forest Service began a long-term soil productivity (LTSP) study in major commercial timber types within national forests across the country. The SRS is studying the loblolly pine type and has set up installations in Texas, Louisiana, Mississippi, and North Carolina. Following the harvest of a mature stand of trees, research scientists and their partners studied nine combinations of soil compaction and organic matter removal. The ecosystems now developing on those sites are monitored closely to determine relationships among soil compaction, organic matter removal, and tree growth. A series of companion studies involving forest industry, the SRS, and several universities have begun. Dubbed MPEQ (Monitoring Productivity and Environmental Quality in Southern Pine Plantations), the effort includes studies in east Texas, north and southeast Louisiana, and South Georgia. In both the LTSP and MPEQ studies, timber stands are documented by extensive sampling of soils and all aboveground vegetation. Pine growth and a variety of soil and other biological processes will be monitored through the next rotation. The regional nature of the LTSP study, and

166 the types of environmental monitoring that are conducted, have enabled scientists from four SRS research work units to begin a study of coarse woody debris decomposition on the sites. Decomposition rates for various sizes of woody debris will be correlated to environmental conditions on the sites and to the role of termites and other wood-inhabiting insects. Over time, changes in chemical composition (tannins and structural chemicals) also will be determined and related to other measured variables.

3. Southeast Tree Research and Education Site (SETRES) SETRES was established in 1992 as a major project of the Southern Global Change Program. It was designed to examine the interactive responses of loblolly pine growth and physiology to changes in atmospheric carbon dioxide (CO2), nutrition, and water. SETRES is a strong and active collaboration of the Forest Service, North Carolina State University’s Forest Nutrition Cooperative, and several industrial partners. CO2 experiments have moved from installing branch bags to enclosing entire 14-year-old trees in open-top chambers. This work was completed in winter 1999. Stand responses are being assessed and, so far, have clearly demonstrated the plasticity of loblolly pine in response to fertilization. Growth rate has tripled in 5 years. In addition, collaborative partners have joined the study and are using the well-executed and maintained experimental design. Collaborative projects like SETRES are good examples of work performed under a crosscutting theme framework. The following are other examples of ongoing research on this important site: Impacts of elevated CO2 on loblolly pine physiology (NC State); whole-tree and stand water relations (Duke University); treatment impacts on wood quality (NC State); CO2 impacts on pest resistance (NC State); impacts of fertilization on soil water quality (Duke University); impacts of treatments on aspects of long-term soil productivity (NC State and Purdue University); root growth and dynamics (Duke); seasonal variation in carbon gain as affected by environmental responses (Virginia Tech); testing and applying mathematical growth models (Southern Global Change, University of New Hampshire, Louisiana State University, Oak Ridge National Laboratory, CSIRO-Australia); and inclusion in a long-term multi-site soil archives (Duke University). Most of these projects included the training of graduate students; and SETRES contributes in large measure to the continuing education of southern foresters. SRS scientists will maintain SETRES as a long-term experiment and include it as a centerpiece of regional studies.

4. Regional Competition Control Project In the early 1980’s, the USDA Forest Service Vegetation Management Project in Auburn, Alabama, began developing a cooperative, long-term study known as the Competition Omission Monitoring Project (COMP). The project is composed of a multi-disciplinary group of cooperators from the former Southern and Southeastern Forest Experiment Stations, several southern universities, and many forest industry cooperators. The group operates under the premise that increases in crop tree growth and yield alone would be insufficient to justify some forest vegetation management treatments. On some ownerships, these benefits would have to be weighed against possible changes in soil productivity, wildlife habitat, biodiversity, and wood quality, as well as the possible effects of treatment on insect populations and pathogens. The study’s key features include its uniform study design and sampling protocol on each of the 13 sites, which are found from Louisiana to Virginia in several physiographic provinces. After more than 15 years of study, several multiple-author papers have been published; and the study remains viable, even though there have been major changes in corporate ownership and investigation staff. The study’s lasting strengths stem from an informal team organization that is committed to sustaining high-value, cooperative, long-term studies; where the most meaningful information is derived only after 10 to 20 or more years of continuing investigation. In addition to addressing the study’s first objectives, COMP also is considering several economic evaluations of commodity and non-commodity values. Its members are conducting surveys of “public preferences” and how they are linked to study treatments. The COMP framework also provides opportunities to examine the ecology of mixed- and single-species stands; and the processes that influence resource conservation, uptake, and cycling within stands that developed on competition-control treatment sites.

The spirit of collaboration represented in each of these four examples is now being extended to explore new research and development opportunities. For example, several SRS research units and their university partners have recently begun collaborative studies to assess the use of fire and other tools for reducing or removing unhealthy forest fuel accumulations. Their efforts are helping to reduce the risk of wildfires and are also helping to restore health and productivity to southern pine ecosystems.

167 Outcomes/Products: We consider development of the Southern Pine crosscutting theme to be a practical endeavor designed to produce tangible products. In addition to filling research gaps, much progress can be made by simply integrating current knowledge. Outcomes from this integrated approach will strive to ensure that:

1. Science-based information is available to all interested users and contributes to forest management practices on mixed land ownerships. 2. Research programs are substantial, integrated, and well organized. They respond to the needs of all forest users, to whom research results are widely disseminated. 3. Management options are designed to maintain forest ecosystem processes, functional relationships, and structure at all spatial levels. 4. Station personnel provide information about the southern pine and pine-hardwood ecosystems, which will enhance a broad range of social, environmental, economic, and cultural values. 5. Cooperation and coordination among and between landowners, agencies, and organizations is used to achieve broad societal goals with regard to the southern pine ecosystem. 6. Improved forest management practices, reduce the number of threatened, endangered, and sensitive (TES) plant and animal species that are listed.

To achieve such results we will need a number of new tools:

1. Integrated models that predict the effects that alternative vegetation management and harvesting treatments would have on plant succession, floral and faunal diversity, soil, water, wildlife, timber growth and properties, ecosystem structure and function, and economic efficiency. 2. Guidelines for managing pine and pine-hardwood forests that will simultaneously meet varied landowner objectives and sustain productive, functional ecosystems. 3. Guidelines for restoring longleaf pine and other ecosystems, including the use of prescribed fire to economically restore native flora and associated fauna. 4. Operational models for mitigating smoke hazards, and documentation of the long-term effects that season and fire frequency have on tree growth, coarse woody debris and snags, and the composition and structure of understory vegetation, including TES species. 5. Documentation of temporal trends in resource conditions and implementation of monitoring to evaluate the influence of management practices on long-term productivity. 6. Documentation of the socio-economic, legal, tax, institutional and demographic effects of alternative management practices, land-use changes, and the associated fragmentation of forest. 7. Guidelines based on cutting-edge science and technology that optimizes timber production on selected ownerships.

There are now opportunities for government, industry, and universities to develop collaborative, cooperatively funded research programs; programs that will help move forest science forward into a new century. We are confident such an approach will produce research results that are both useful to domestic and international forest policymakers and beneficial to those they serve.

LITERATURE CITED Burkett, V., S. Beasley, P. Roussopoulos and J. Barnett. 1996. Toward southern forest sustainability: a science agenda. Seventh American Forest Congress, Southern Region Forest Research Committee Report. Seventh American Forest Congress, Washington, DC. 27 p. Canadian Forest Service. 1995. Criteria and indicators for the conservation and sustainable management of temperate and boreal forests; the Montreal process. Canadian Forest Service, Natural Resources Canada, Quebec, Canada.27 p. USDA Forest Service. 1997. Strategic framework for the Southern Research Station. USDA Forest Service, Southern Research Station, Asheville, NC.24 p. USDA Forest Service. 2000. Sustainability and Productivity of Southern Forest Ecosystems. USDA Forest Service, Southern Research Station, Science Update, SRS-001, Asheville, NC.27p. World Commission on Environment and Development. 1987. Our common future. Oxford, United Kingdom: Oxford University Press.

168 LONGLEAF PINE: AN UPDATED BIBLIOGRAPHY - PART II 1995-2000

Charles K. McMahon (USDA Forest Service, Southern Research Station, 520 Devall Dr, Auburn, AL 36849) John S. Kush (Auburn University, School of Forestry & Wildlife Sciences, Auburn University, AL 36849)

ABSTRACT: In early 1994 the USDA Forest Service, Southern Research Station initiated a five-year program known as the Longleaf Pine Ecosystem Restoration and Management Program. This program was aimed at conducting collaborative research studies and building partnerships that would help reverse the decline of this critical Southern forest ecosystem. As part of the initial program activities, an updated Longleaf Pine bibliography was commissioned in cooperation with Auburn University School of Forestry and Wildlife Sciences. An update was compiled containing three previous bibliographies (Wahlenberg 1946, Croker 1968, and Hu and Burns 1986) along with 500 new entries. A hardcopy version containing the 500 new entries (up to 1995) was published in 1996. In addition, an electronic database containing both old and new entries was made available over the Internet. The plan was to update the database approximately every 5 years. This latest update contains many new entries, primarily from the period 1995-2000. The bibliography is available for searches and/or downloading at no cost from the Longleaf Alliance web site at Auburn University, http://www.forestry.auburn.edu/la. Suggestions for additional entries are always welcomed and should be e-mailed to [email protected].

_____

In early 1994 the USDA Forest Service, Southern Research Station initiated a five-year program known as the Longleaf Pine Ecosystem Restoration and Management Program. This program was aimed at conducting collaborative research studies and building partnerships that would help reverse the decline of this critical Southern forest ecosystem. As part of the initial program activities, an updated Longleaf Pine bibliography was commissioned in cooperation with Auburn University School of Forestry and Wildlife Sciences. An update was compiled which contained the three previous bibliographies (Wahlenberg 1946, Croker 1968, and Hu and Burns 1986) along with 500 new entries. A hardcopy version of the bibliography containing the new entries (up to 1995) was published in 1996 (Kush et al, 1996). In addition, an electronic database containing the older bibliographies and the new entries was compiled and made available on-line and linked to the Longleaf Alliance web page. The plan was to update the database approximately every 5 years.

This latest update (part II 1995-2000) contains 531 new entries. This doubling of entries over the past 5 years is somewhat expected given the high recent interest related to the restoration and management of longleaf pine ecosystems. This interest resides in all of the states in the longleaf pine historical range and exists on both public and private ownerships. The driving force of this interest is connected to the ecological, economic and cultural values embedded in the tree and the surrounding plant and wildlife communities. As a result, we now see a significant number of “longleaf pine” related publications in many new outlets. A much wider array of “longleaf” topics are now under investigation and being published as compared to Wahlenberg’s time. We expect this trend to continue over the next five years.

The bibliography with the current update is still available for searches and/or downloading at no cost from the Longleaf Alliance web site at Auburn University, http://www.forestry.auburn.edu/la. Suggestions for additional entries are always welcomed and should be e-mailed to [email protected].

169 STATUS OF RECENT LONGLEAF PINE REGENERATION ON EGLIN AIR FORCE BASE (OR A SANDHILLS SITE IN NORTHERN FLORIDA)

Ralph S. Meldahl (Auburn University School of Forestry & Wildlife Science, 108 M. White Smith Hall, Auburn University, AL 36849) John S. Kush (Auburn University School of Forestry & Wildlife Science, 108 M. White Smith Hall, Auburn University, AL 36849) J. Morgan Varner (Interdisciplinary Ecology Program, Box 118526, University of Florida, Gainesville, FL 32611-8526)

ABSTRACT: Natural regeneration of longleaf pine is one of the most important tools natural resource managers have at their disposal to regenerate existing longleaf pine stands in the southern United States. However, adequate cone crops for natural regeneration typically occur every 5-7 years, and often longer. The 1996 longleaf pine seed crop was the largest in the Florida Panhandle in more than 30 years. In 1997, a study was established on sandhill sites on Eglin Air Force Base in northern Florida to monitor longleaf pine seedlings from the 1996 seed crop and advanced regeneration in relation to various gap sizes. The overstory longleaf pines surrounding each gap were stem mapped. Gap sizes varied from 0.10 to 1.70 acres. Seedlings and advanced regeneration are being monitored for survival, root-collar growth, and time to height initiation. Four transects were established in each gap, and the 1996 seedlings and ground cover were stem mapped within. Recruitment ranged from 31,950 to 223,289 seedlings/hectare. Transects have been re-inventoried each year since. First-year survival was 38.3% and two-year 24.2%. Among advanced regeneration still in the grass-stage, nearly 50% of the seedlings initiated height growth between 1997 and 1999, with mortality less than 4%. Ground-line diameters for these seedlings increased nearly 72% between 1997 and 1999. Survival for 1996 seedlings appears to be associated with ground cover on these droughty sites. Seedlings in close proximity to ground cover had increased survival compared to seedlings with no ground cover nearby.

INTRODUCTION Longleaf pine forests are critically endangered communities of the southeastern USA. Aside from being the dominant tree species, longleaf pine is also considered a keystone species of this community type. Reestablishing and regenerating longleaf pine seedlings is a major priority for conserving and restoring longleaf pine forests throughout the region. This fact is especially true for the sandhills communities of northern Florida, USA, an area with historically low success in longleaf pine regeneration, resulting in poorly stocked stands.

The largest communities, area-wise, of longleaf pine are in public ownership, including National Forests, National Wildlife Refuges, and Department of Defense Installations. Eglin Air Force Base (Eglin), located in northwestern Florida, USA, represents the largest remaining single acreage of longleaf pine forest. Eglin is the largest forested military reservation in the free world, consisting of approximately 188,000 ha. A sandhills ecological association covers 78% of Eglin that is characterized by rolling sandhill ridges dissected by streams.

Natural regeneration is the most important tool natural resource managers have at their disposal to ecologically and economically regenerate existing longleaf pine stands. However, adequate cone crops for natural regeneration typically occur every 5-7 years, and often longer (Wahlenberg, 1946; Maki 1952).

PROBLEM? Land Use History & Present Management Prior to World War II, Eglin was managed by the USDA Forest Service as the Choctawhatchee National Forest. As part of their management, a majority of the sandhills were selectively logged and heavily turpentined. Early US Forest Service management left 6 to 10 seed trees ha-1 for natural regeneration purposes.

Lack of fire as a management tool allowed scrub oaks to take over longleaf sites, shading out ground cover plants and maintaining poor stocking. Today, these xeric sandhill sites are dominated by an overstory of scattered longleaf pine (Pinus palustris), with an understory of turkey oak (Quercus laevis), bluejack oak (Q.

170 incana), and sand post oak (Q. margaretta). The ground cover consists of various grasses, herbs, and runner live oak.

ANSWER! Today, one of the primary management goals on Eglin is restoration and maintenance of longleaf pine communities on sandhill sites. One of the major tools Eglin uses to accomplish this goal is the utilization of natural regeneration techniques. The 1996 longleaf cone crop represented a unique opportunity to obtain longleaf pine natural regeneration wherever needed. This is especially important at those locations where good longleaf cone crops are rare.

RESEARCH OBJECTIVES To develop a monitoring program to examine relationships among the 1996 longleaf pine seed crop, recruitment in gaps, seedling/sapling mortality, and subsequent growth.

METHODS A. Measure canopy gaps. We located and measured length, width, and orientation of variably sized gaps. B. Monitor the 1996 seedling crop: We established four transects per gap, extending from 1.8 m from the gap center to a point at least 4 m beyond the gap edge. Transects were installed along the long axis of the gap and its right angle. Transects are 45 cm wide and seedlings are mapped on an X-Y basis (photograph below). Vegetation (Woody, Herbaceous, Litter, or Bare Ground) along the transect was surveyed. C. Stem map seedlings established prior to 1996 seed crop: Longleaf pine seedlings have the unique growth stage termed the “grass stage”, where height growth is ignored in favor of taproot development. Height growth initiation occurs when seedlings attain a ground line diameter of approximately 2.5 cm. In order to predict future survival, growth, and height growth initiation of the 1996 seedling crop, we tagged 25 existing grass stage seedlings per plot (photograph below) and measured distance, and direction from plot center, and ground-line diameter (GLD). The year of these seedlings’ establishment was unknown, but we suspect that they are the result of the last successful seed crop at Eglin, in 1987. D. Stem map the surrounding longleaf pine overstory: Record diameter at breast height (1.37 m; DBH), crown class, and crown and total height (Figure 4).

RESULTS A. Canopy gaps. Selected gap sizes averaged 0.22 ha (2200 m2) with a range from 0.04 to 0.69 ha (400 to 6900 m2); 80% of gaps observed were elliptical in shape; 44% of the gaps were oriented in an east-west direction.

B. 1996 Seedling Crop. Initial Seedling Stocking (1997) Mean seedlings per gap: 120,045 seedling ha-1; Range: 31,950 to 223,289 seedlings ha-1 No significant relationship between initial stocking and gap size PROC GLM Prob. > F 0.4424 No significant differences between initial stocking and axis aspect North (316 to 45 degrees): 114,398 East (45 to 135 degrees): 123,454 South (136 to 225 degrees): 129,612 West (226 to 315 degrees): 112,716 t-test Prob. > F 0.8880; No significant differences between stocking and vegetation type.

First to Third-Year Seedling Survival (1998, 1999 & 2000) First year (1998) survival was 38.3% Second year (1999) survival decreased to 24.2% Third year (2000) survival decreased to 5.8% Survival was not significantly associated with gap size PROC GLM Prob. > F 0.0943 Survival was not significantly associated with axis aspect North (316 to 45 degrees): 35.4% East (45 to 135 degrees): 35.4% South (136 to 225 degrees): 38.2 %

171 West (226 to 315 degrees): 44.3% Duncan’s test. Prob. > F 0.4366; overall mean 38.3%

Survival at the interior of the gap (nearest 15.2 m to gap center) was not significantly different from survival at the edge of the gap (nearest 15.2 m from gap edge). Survival was significantly associated with vegetation cover as seedlings associated with woody cover had a higher survival percentage than seedlings associated with herbaceous and litter cover or bare ground.

C. Pre-1996 Seedling Growth and Survival. In 1997, 272 grass stage longleaf pine seedlings were tagged and mapped. Average 1997 GLD was 1.63 cm with a range of 0.284 to 3.75 cm. In 2000, average GLD increased to 3.79 cm with a range of 0.895 to 6.56 cm. GLD increased 81.2% between 1997 and 2000. There was only 3.7% mortality since 1997, with a range of 0% to 20.0% per gap. Nearly 65% of the seedlings initiated height growth between 1997 and 2000.

DISCUSSION Data collected from 3 years of studying the relationship between longleaf seedlings and gap size does not agree with other published data. Brockway and Outcalt (1998) found significantly fewer seedlings within 12m of adult trees growing along the gap edge. Boyer (1993) reported adult longleaf pines had a zone of seedling exclusion and this zone extended farther on poorer sites. This study indicates there is no significant difference between survival near adult trees and survival in the gap’s center.

Although preliminary, our results are counter to the prediction that sunnier exposures (north and east locations) have better growth and survival than shadier exposures (west and south). Our data tend to show the opposite, with the shadier exposures and shadiest vegetative cover (woody) having better survival. 47.0% of trees > 23 cm (seed producing diameter) Average total height of tress > 23 cm was 16.6 m; average crown height was 10.5 m

FUTURE RESEARCH EFFORTS A. A growing season fire will be imposed on at least one location to study its effects on the survival of the 1996 seed crop. Plans were to incorporate this into the study in 1997 but dry conditions during the past 3 growing seasons have prevented the introduction of fire. B. The 1996 seed crop’s annual survival will be monitored for at least 1 more year. Of special interest will be the impact of fire on the vegetation cover and its impact on seedling survival. C. The pre-1996 seedlings growth and survival will continue to be monitored. The effects of fire on growth and survival will also be monitored. D. The surrounding overstory trees’ and stand’s age and recent radial growth will be assessed during the dormant season.

DESIRED FUTURE CONDITIONS The photographs below represent the conditions the natural resource managers at Eglin Air Force Base are trying to attain. With longleaf pine cone crops large enough to be considered adequate for natural regeneration as infrequent as every 25 years, they need to be able to protect what seedlings they have to help them emerge from the grass stage. Only by understanding gap and replacement patterns can managers restore the longleaf pine forest ecosystem.

ACKNOWLEDGEMENTS Funding for this project has been provided by the US Forest Service and the Natural Resources Division of Eglin Air Force Base. This project would not have been possible without the field and laboratory efforts of Dick Sampson, Chadwick Avery, Bryan Lindsey, Wes Hauffe, Marie Earle, Shawn Harrison, Dennis Shaw, Eric Reynolds, Anshu Shrestha, Lee Helton, and Jessica Cochrane. Steve Seiber, Scott Hassell, and Tim Christiansen at Eglin Air Force Base have provided assistance while working on the Base.

172 LITERATURE CITED Boyer, W.D. 1993. Long-term development of regeneration under longleaf pine seedtree and shelterwood stands. South Journal of Applied Forestry 17(1):10-15. Boyer, W.D., and J.B. White. 1990. Natural regeneration of longleaf pine. In: Proceedings of the symposium on the management of longleaf pine, R.M. Farrar, Jr. (ed.), USDA Forest Service, General Technical Report SO-75, pg. 94-113. Brockway, D.G., and K.W. Outcalt. 1998. Gap-phase regeneration in longleaf pine wiregrass ecosystems. Forest Ecology and Management 106:125-139. Maki, T.E. 1952. Local longleaf seed years. Journal of Forestry 50(4):321-322. Wahlenberg, W.G. 1946. Longleaf pine: Its use, ecology, regeneration, protection, growth, and management. Charles Lathrop Pack Forestry Foundation in cooperation with the USDA Forest Service, 429 p.

173 OVERSTORY STRUCTURE AND REGENERATION PROCESSES IN LONGLEAF PINE- WIREGRASS FORESTS

Robert J. Mitchell (Joseph W. Jones Ecological Research Center, Newton, GA) Brian J. Palik (USDA Forest Service, Forestry Sciences Laboratory, Grand Rapids, MN) Robert H. Jones (Department of Biology, Virginia Polytechnic Institute and State University, Blacksburg, VA) Mou Pu (Department of Forestry, Virginia Polytechnic Institute and State University, Blacksburg, VA) Stephen D. Pecot (Joseph W. Jones Ecological Research Center, Newton, GA) Stacy Hurst (Joseph W. Jones Ecological Research Center, Newton, GA) Michael A. Battaglia (Department of Forestry, Virginia Polytechnic Institute and State University, Blacksburg, VA) Glen L. Stevens (Department of Biology, Virginia Polytechnic Institute and State University, Blacksburg, VA)

ABSTRACT: Silvicultural methods used to meet objectives of ecosystem management often include green tree retention, or a reserve shelterwood, in order to maintain components of mature stand structure. The competitive environments and mechanisms that influence regeneration in such systems differ substantially from those under even-aged management. We initiated a study in a 65-year-old longleaf pine forest to address the effects of residual overstory structure and competing herbaceous vegetation on survival and growth of longleaf pine seedlings. Stands were harvested to similar residual basal areas using single-tree selection, small group (~0.25 ac) selection, and large group (~0.5 ac) selection. An uncut control stand was used as a reference. Twenty-five subplots encompassing the range of overstory abundance index (OAI) were installed. Ten one-year-old containerized longleaf pine seedlings were planted at 2 quadrats (2 by 2-ft spacing) per subplot, one side receiving a glyphosate application to remove the understory. Soil resources and light availability were quantified over a two-year period, as well as seedling survival, size, and growth. Trench plots were also installed across a range of OAI.

The overstory and understory facilitated survival of longleaf pine seedlings but competed with them relative to seedling growth. Seedling survival increased positively with OAI and in the presence of understory. Microclimate changes, i.e., lower soil temperature and relative humidity, in the shade appeared to result in facilitation rather than increased soil moisture during a severe drought. Seedling growth was negatively influenced by OAI largely through attenuation of light by the understory. Soil N was increased at low OAI but only when the understory was absent. Root gaps created through overstory removal appear to be filled quickly by understory plant communities. Finally, trenched subplots preventing overstory root encroachment resulted in a substantial hardwood response which negatively affected growth and survival of longleaf pine seedlings.

174 THE LONGLEAF PINE GROWER’S GUIDE TO SUCCESS: A GUIDE FOR PRIVATE LANDOWNERS

Julie H. Moore (1918 Fuller Street, Hattiesburg, MS 39401) Lawrence S. Earley (126 Forest Road, Raleigh, NC 27605)

ABSTRACT: This guide provides the private landowner with information on successful ways to manage longleaf pine woodlands and maintain the natural integrity of the ecosystem while obtaining economic return. Region wide, the private sector controls over 51 percent of the remaining 2.9 million acres of the longleaf pine ecosystem. Owners of longleaf woodlands have a unique opportunity to make a tangible contribution to the future of this valuable, now unusual, forest system. This 32-page booklet illustrated with photographs and drawings, features interviews with six landowners of longleaf woodlands from Texas to North Carolina and describes a variety of management objectives, challenges, and successes. Benefits derived from promoting longleaf pine forests are discussed, including timber, pine straw, wildlife habitat enhancement and hunting, aesthetics and our southern heritage. The forces, particularly fire, that have created the longleaf ecosystem are summarized and the four common landscapes supporting longleaf are described and illustrated. Harvest and regeneration methods that approximate natural processes are described and illustrated.

The Nature Conservancy believes that conservation management is crucial to the future of longleaf. While ecologists continue to discover and assemble pieces of the longleaf pine ecosystem puzzle, landowners can harvest an array of valuable products it produces. With an understanding of the basic forces that created the longleaf ecosystem (fire, weather, soil types/moisture and fuel) and by mimicking natural processes during harvesting and regeneration, forest growers can remove a variety of forest products in a sustainable manner with minimal resulting impact to all the components of a healthy forest.

Recognizing the need for a non-technical publication that provides management suggestions based on fundamental ecological concepts, The Nature Conservancy provided the initial funding for this guide. Belief in the superiority of longleaf as a renewable timber type and the associated values for landowners is a basic premise of The Longleaf Alliance, the other sponsor of this publication. “MANAGING THE FOREST AND THE TREES” will be available from state field offices of The Nature Conservancy and from the Longleaf Alliance in January 2001.

CONTENTS * The Forest is More Than the Trees: the longleaf pine ecosystem, the role of fire and the four basic landscape types * Ecological Management of Longleaf Pine: Debunking the myths and frequently asked questions about managing for longleaf * Six Case Histories: Landowners talk about how they manage longleaf * Lessons of Longleaf Pine Management: Ten characteristics of productive native forests and how to achieve and maintain them * Where to Get Help: Sources of technical information, assistance with developing management plans and cost-sharing and other financial assistance * Where to See Longleaf Pine: From Virginia to Texas where to see ongoing management and restoration in progress * Questions to ask Foresters and Loggers Before Hiring Them * What to Include in a Harvest Contract

175 CLASSIFICATION OF LONGLEAF PINE COMMUNITIES IN THE UPLAND ISLAND WILDERNESS AREA, TEXAS

Brian P. Oswald (Arthur Temple College of Forestry, Stephen F. Austin State University) Leslie Dale (Arthur Temple College of Forestry, Stephen F. Austin State University) Rick Turner (The Nature Conservancy, Nacogdoches, Texas)

ABSTRACT: Native longleaf pine forests were known for their biodiversity, often having the greatest vegetative diversity on any North American forest. This diversity, encouraged and developed through native anthropogenic activities, was greatly impacted by Euro-anthropogenic management practices. Restoration efforts of longleaf pine often center attention on this biodiversity issue. What has happened to this biodiversity on lands where little or no management activities have been applied for a number of decades?

Eighteen plots were established within the naturally established longleaf pine-dominated stands of the Upland Island Wilderness in East Texas to evaluate the vegetative diversity within areas where no active management activities have occurred in 20 years. The species and number of stems of the overstory, mid- story and understory layers of these stands were measured. A variety of multivariate statistical tests (DCA, Cluster Analysis) were applied to determine what type of differences were found as well as where differences in diversity occurred. Variation between stands were found within each vegetative layer. Specific results and possible rationale for these differences are discussed.

_____

What has happened to the biodiversity of longleaf pine ecosystems on lands where little or no management activities have been applied for a number of decades? Are natural disturbances such as fire and inherent differences in site characteristics capable of maintaining different longleaf pine communities twenty years after active management activities have ceased? The objective of this study was to classify longleaf pine communities within a small area in East Texas, and attempt to determine what important environmental gradients may explain any differences among these communities.

As part of a classification project of two wilderness areas in East Texas, 107 plots were established within the Upland Island Wilderness Area. Initial stand delineation was completed through the interpretation of leaf-on color infrared aerial photographs. Topographic maps were then used to further separate mixed-species stands based on topographic features. A dot grid was used to randomly locate sapling plots within each identified stand. Out of the 107 plots thus identified, 19 were placed within naturally established longleaf-pine dominated communities.

Sampling of the overstory species was performed within a 0.04-hectare plot. Midstory (shrub and sapling) species were measured within a nested 0.01-hectare plot. Two, 1m2 plots were located at opposing corners of the overstory plot to measure understory (herbaceous and woody seedling) species. Detrended Correspondence Analysis (DCA) was used to perform a series of ordinations of the sample plots based on each vegetative stratum (overstory, midstory, understory) both separately and in combination. TWINSPAN cluster analysis was then used to classify the plots based on species composition.

Differences in overstory composition appear to be related to a soil moisture gradient and fire frequency. Some plots were found on more xeric sites, or those with signs of recent fire history (such as fire scars, top- killed hardwoods, etc.), while others were more mesic, supported more hardwoods, and showed few, if any signs, of past fire occurrence.

Differences in the midstory were also apparently related to soil moisture and fire occurrence. The hardwood component (# species, # individuals/species) was drastically reduced in the four plots with the greatest signs of past fire. Understory composition appears to be driven by soil moisture and canopy closure. Two plots were found adjacent to an old beetle spot in which overstory pines were killed, and in a very open longleaf stand, with little midstory.

176 When all of the vegetative strata were combined, two distinct groups may be observed. Six plots are those with few if any hardwoods in the strata, are more open, and have more recent fire histories, with more herbaceous vegetation. We classified these as Longleaf Pine Woodlands. The remaining plots were classified as Mixed Longleaf Pine-Hardwood Forests.

Soil moisture and fire frequency appear to be the main factors influencing species composition within the longleaf pine communities of Upland Island Wilderness. Without widespread fire occurrence, succession is proceeding within these communities to a mixed longleaf pine-hardwood community. These communities have higher canopy coverage, and support less grass and other herbaceous species, but more shrub species and a thicker duff layer. There is also less longleaf pine regeneration in these communities.

The development of a thick shrub and midstory vegetative layer creates a different fuel complex in these communities. The increase in ladder fuels increases the chance of a stand-replacing fire rather than the more natural low-intensity surface fire usually associated with longleaf pine communities. If such stand-replacing fires occur, the resulting post-fire forest community will likely have a very different species composition than the communities currently found in Upland Island. Since the probability of naturally occurring fires is low in small wilderness areas such as Upland Island, human-ignited prescribed fires may be necessary to maintain the community structure we associate with longleaf pine ecosystems.

177 CONDITION OF LONGLEAF COMMUNITIES IN THE SOUTHEAST

Kenneth W. Outcalt (USDA Forest Service, Athens, GA 30602)

ABSTRACT: The objective of this study was to determine the condition of the remaining longleaf habitat and to document relationships between that condition and location, ownership and burning history. Naturally regenerated longleaf pine stands were randomly selected for sampling from all the stands dominated by longleaf on the Forest Inventory and Analysis (FIA) plots maintained by the U.S. Department of Agriculture, Forest Service. Most stands had a mixture of different size classes with 70 to 80 percent containing at least three sizes and over 50 percent containing trees of all sizes (seedling, sapling, pole, and sawtimber). Nearly all stands had either sapling or seedling size regeneration. Older trees were not very prevalent in Florida or Georgia, where about 10 percent of the stands contained longleaf older than 90 years. North Carolina had the greatest percent of stands with trees larger than 50 cm diameter, while Florida had the least. Seventy percent or more of the longleaf understory on public owned lands was rated as fair to good condition, while less than half of private lands were rated fair or above. Significant disturbance to the soil and understory was most prevalent on private lands and quite rare on public lands. Recent use of fire was found on most public lands, but fire was much less common in private held longleaf stands. Because of soil disturbance and infrequent burning, many private lands containing longleaf are likely to suffer further habitat degradation. The potential exists also for significant loss, since most of the longleaf stands have trees in the sawtimber and pole size class that are vulnerable to immediate harvest.

INTRODUCTION Longleaf pine ecosystems once occupied as much as 25 million hectares in the southeastern United States, extending south from southeast Virginia to central Florida and west into eastern Texas (Stout and Marion 1993). Longleaf pine was native to a wide range of ecosystems including wet flatwoods and savannas along the Atlantic and Gulf coastal plain and higher droughty sand deposits from the fall line sandhills to the central ridge of Florida (Stout and Marion 1993). Longleaf pine even extended onto the mountain slopes and ridges of Alabama and northwest Georgia, where it was found growing at elevations up to 600m (Boyer 1990). Logging of the valuable longleaf pine forests began in colonial times, reached a peak shortly after 1900 (Ware and others 1993). Clearing of forestland for urban and agricultural uses, conversion to loblolly and slash pine plantations, and harvest without regeneration all contributed to the continuous decline of the once dominant forest type of the south. By 1935, only about 8.1 million hectares of longleaf pine forest remained. The amount declined to 4.9 million hectares in 1955, 1.5 million hectares in 1985 (Kelly and Bechtold 1990), and 1.2 million hectares in 1995 (Outcalt and Sheffield 1996).

Prior to landscape fragmentation brought by human habitation, fire was a frequent, natural occurrence across much of the Southeast maintaining extensive longleaf pine and grass communities. Without fire, plant community composition and structure changes, as woody species increase and grasses and forbs decline. It is widely recognized that loss of longleaf habitat plus decline in health of remaining areas due to lack of fire, would lead to widespread endangerment of the myriad of species that use the longleaf ecosystem. The objective of this study was to determine the condition of the remaining longleaf habitat in the southeast and to document relationships between that condition and location, ownership and burning history. This information should be useful in developing a strategy for maintaining and/or restoring longleaf habitat in this region.

METHODS All sampling was based on the network of FIA plots in Florida, Georgia, South Carolina, and North Carolina, which have been established by the U.S. Department of Agriculture, Forest Service, Southern Research Station. To obtain a proportionate sample of all major forest types, sites, and ownership classes in each state, scientists systematically distributed permanent plots across all ownerships. For this study, all plots in the selected states with a longleaf-dominated overstory that resulted from natural regeneration were sorted into 5 age groups: 1–10, 11–20, 21–40, 41–60, and 60+ years. Age groups were classified by soils and topography into site types, which varied from state to state but were grouped into wet lowlands, dry sandhills, and rolling uplands.

178 Within each age class and site type, 3 stands were randomly selected for sampling. In Florida, we sampled 45 stands while 30 stands each were sampled in Georgia, South Carolina, and North Carolina. Tree density and size was sampled in 10 randomly located circular plots in each stand. We also recorded any evidence of past soil disturbance or recent fire. The condition of the understory community was determined by general appearance, dominance of typical native species, and amount of human disturbance. Data were summarized and percents calculated for each state by ownership class, i.e., public or private.

RESULTS A number of the stands were converted to other uses before we could sample them. The losses were highest in Georgia where 25% of the selected stands were cleared for agricultural or urban uses. About half of all sample stands had more than 25 sawtimber sized (>22 cm diameter) trees per hectare while 75% had at least some sawtimber on them. The majority of stands were multisized or all aged with about 50% containing longleaf from seedling to sawtimber size in Florida and Georgia and about 70% in North and South Carolina. Natural regeneration was present in at least 90% of the sampled stands as either seedlings or saplings (<10 cm diameter). North Carolina had the greatest proportion of stands with mature trees (>90 years) at 28% while Georgia (10%) and Florida (12%) had the lowest percentage of stands with old longleaf trees. Trees larger than 50 cm diameter were found in over 20% of longleaf stands in Georgia and North Carolina, but were very rare in Florida (2%).

The majority (80%) of sampled longleaf stands were in forested areas with only a small fraction surrounded by non-forest agricultural or urban landuses. Georgia longleaf stands had fewer recent human caused disturbances (20%) than the other states where about 40% of the sampled stands had moderate to severe disturbance to soils and vegetation. This disturbance was much more prevalent on private lands where it ranged from a low of 25% in Georgia to a high of 75% in Florida. Less than 15% of public lands had evidence of recent soil or vegetation disturbances other than fire. Fire, however, was quite common on public lands where over 80% had been burned within the last 5 years. Burning of privately owned longleaf stands was highest in Georgia (48%), intermediate in Florida (35%) and South Carolina (30%), and lowest in North Carolina (15%). Average shrub height was influenced by burn history with a mean height of 3.0m in unburned stands and 2.0m in those recently burned. The majority of public longleaf stands had an understory in good or fair condition while many private longleaf stands were judged as poor or very poor in understory health.

CONCLUSIONS Natural stands of longleaf continue to disappear in all four southeastern states due to conversion to other species or land uses. Losses appeared highest in Georgia because the survey used for selecting stands was 8 years old while others had been done in the last 3 to 4 years. Those natural stands of longleaf that remain are a mixture of size and age classes. Most longleaf stands have regeneration, indicating that recruitment is occurring to replace trees that die. Mature longleaf pine, those greater than 90 years old, are not very prevalent, which means the old growth that does exist has a high ecological value. Most of the remaining stands are in forested settings, which should make application of the prescribed burning needed to maintain ecosystem health easier. Burning, however, is not being used on many of the privately owned longleaf stands, while they are receiving significant amounts of degrading soil disturbance. In addition, many contain trees in the pole and sawtimber size class, which make them vulnerable to immediate harvest. All these factors stress the need for a continued effort to provide information and assistance to landowners to help them make informed decisions.

LITERATURE CITED Boyer WD. 1990. Longleaf pine. In: Burns RM, Honkala BH, tech. coords. Silvics of North America, Volume 1 Conifers. USDA, Forest Service, Washington, DC, Agriculture Handbook 654:405-412. Kelly, J.F., and W.A. Bechtold. 1990. The longleaf pine resource. Pages 11-22 in R.M. Farrar, Jr. (ed.). Proceedings of the Symposium on the management of longleaf pine. General Technical Report SO-75, U.S. Department of Agriculture, Forest Service, Southern Forest Experiment Station, New Orleans. Outcalt, K.W., and R.M. Sheffield. 1996. The longleaf pine forest: trends and current conditions. Resource Bulletin SRS-9, U.S. Department of Agriculture, Forest Service, Southern Research Station, Asheville, NC.

179 Stout, I.J., and W.R. Marion. 1993. Pine flatwoods and xeric pine forests of the southern (lower) coastal plain. Pages 373-446 in W.H. Martin, S.G. Boyce, and A.C. Echternacht (eds.). Biodiversity of the southeastern United States: lowland terrestrial communities. John Wiley and Sons, New York. Ware, S., C. Frost, and P.D. Doerr. 1993. Southern mixed hardwood forest: The former longleaf pine forest. Pages 447-493 in W.H. Martin, S.G. Boyce, and A.C. Echternacht (eds.). Biodiversity of the southeastern United States: lowland terrestrial communities. John Wiley and Sons, New York.

180 PREDICTING LONGLEAF PINE STRAW PRODUCTION

Jyoti N. Rayamajhi (11654 Seville Road, Fishers, IN 46038) John S. Kush (School of Forestry and Wildlife Sciences, 108 M. White Smith, Auburn University, AL 36849-5418)

ABSTRACT: There is a prevailing view that longleaf pine cannot economically compete with either loblolly or slash pine. However, the wood products produced by longleaf should make landowners re-think their options. Longleaf pines can produce quality wood products when grown in a variety of densities. Landowners should think about growing and marketing poles and pilings that historically have stumpage prices ranging from 30% to 50% or more over sawlog prices. Another forest product, which can be produced concurrently with poles and pilings, is pine straw. Longleaf straw is in high demand as a landscaping material. Retail prices can vary from around $2.50 to $8.00 per bale. Landowners typically receive from $0.30 to $1.10 per bale. Annual yields of up to 200 bales per acre are possible with actual straw yields depending on stocking, site quality, and age. As part of several studies, leaf litter has been monitored on the USDA Forest Service's Escambia Experimental Forest located in southern AL, just south of Brewton. Three to four 9.6 square foot litter traps were set up on 39 research plots during the summer of 1992 to cover a range of density, age and site quality's. Pine straw has been collected from these traps on a four- to six-week basis since. The needles are oven-dried and weighed for a dry-weight per acre basis. Using the last 7 years of data, number of bales of pine straw produced on a per acre basis will be presented.

INTRODUCTION Pine trees deposit a blanket of needles, often-called pine straw, on the forest floor annually. Pine straw is valued by homeowners, nurseries, and landscapers as a source of mulch. Many forest owners do not realize that it is possible to sell this straw, but in fact wise management of this resource can substantially increase the owner’s income from forestland. It can be an annual to biennial source of income that may exceed that from timber.

Several variables control pine straw yields, including vigor and age of trees, their basal area, the season, and the time interval between rakings. A 15-year old stand being raked for the first time can yield as many as 200 bales per acre. Vigorous, young to middle-aged stands will yield more straw than older, low vigor stands. A low annual yield is 50 bales per acre, a good average is 70 bales, and 100 bales is above average.

METHODS The study utilized 56 permanent 1/5-acre measurement plots located on the Escambia Experimental Forest. All plots are located in naturally regenerated stands that are part of the USDA Forest Service’s Regional Longleaf Pine Growth Study. This 35-year old study has been examining stand dynamics of naturally regenerated stands of longleaf pine across the Southeast.

Plot selection for the study is based upon a rectangular distribution of cells formed by stand age classes ranging from 20 to 120 years, site-index classes ranging from 50 to 90 feet at 50 years, and density classes ranging from 30 to 180 square feet per acre.

At the time of establishment, plots are assigned a target basal area class of 30, 60, 90, 120, 150 square feet/acre or left to grow. They are left unthinned to grow into that class if they are initially below the target basal area. In subsequent re-measurements, the plot is thinned back to the previously assigned target if the plot basal area has grown 7.5 square feet/acre or more beyond the target basal area. The thinnings are generally of low intensity and from below.

RESULTS The figures below (Figures 1-4) present the results of modeling the monthly litter data. Stand attributes that were significant in the model were stand age, stand basal area, stand density, and year. The R2 for the model, however, was only 44%. Site index of the stand was not a significant variable. All of the weights are in oven-dry pounds.

181 Pine straw production based on basal area

5000 4608 4716 4606 4500 4558 4000 4026 3579 3500 3112 3000 2500 2000 1992

Pounds/acre 1500 1000 786 500 0 20 40 60 80 100 120 140 160 180 Basal area class (square feet/acre)

Figure 1. The annual pine straw produced on the Escambia Experimental Forest for the years 1993 through 1998. In addition to the role climate plays in litter production, stand management is a critical factor. The forest is managed under a 3-year prescribed burning regime. Even with no visible scorch, it has been observed that trees will lose needles after a prescribed fire.

Annual pine straw production for 1993-1998

5000 4941 4384 4163 4000 3596 3000 2906 2905 2000

Pounds/acre 1000 0 1993 1994 1995 1996 1997 1998 Year

Figure 2. The annual pine straw produced on the Escambia Experimental Forest based on basal area classes. Pine straw production appears to level off at a basal area of 100-120 square feet/acre. The higher the stand density, the smaller the tree crowns and therefore fewer needles on the trees available for pine straw.

182

Pine straw production based on age class

6000 5089 5000 3991 4000 3574 3417 2976 3000 2688 2000

Pounds/acre 1000 0 20 40 60 80 100 120 Age class

Figure 3. The annual pine straw produced on the Escambia Experimental Forest based on age classes. Younger stands produce more pine straw. It may difficult to determine the age of maximum production because most of these stands have been thinned, reducing stand density and therefore reducing the number of trees providing pine straw.

Pine straw production based on site index class

3900 3823 3800 3700 3700 3600 3500 3400 3407

Pounds/acre 3300 3200 3100 60 70 80 Site index class (height at 50 years old)

Figure 4. The annual pine straw produced on the Escambia Experimental Forest based on site index classes. There is a limited range of site index classes. Most stands on the Forest are in the range of 65-75 feet at 50 years. Conventional wisdom indicates the higher the site index the more pine straw produced.

SUMMARY If a bale of pine straw weighs 25 pounds (data from North Carolina Extension Service, a 20-year old, naturally regenerated longleaf pine stand could produce as much as 204 bales/year on an average site. At $0.70/bale, a landowner could make $142/acre/year without doing anything. This is income that could be generated every few years depending on management objectives.

183 THE FRANCIS MARION NATIONAL FOREST 11 YEARS AFTER HURRICANE HUGO

Robin Roecker (Francis Marion & Sumter National Forests, 4931 Broad River Rd., Columbia, SC 29212)

ABSTRACT: The Francis Marion National Forest, comprising 251,700 acres on the Atlantic Coastal Plain of South Carolina, is home to some of the largest tracks of longleaf pine in the southeast. Once dominanted by longleaf pine, much of the forest was cleared in the early 1900’s and began reseeding naturally to loblolly pine. Forest historian Thomas Clark described southern pine lands, in 1925, as “a blackened fire-scorched world, dominated by millions of stumps”. In 1939, the Francis Marion National Forest was purchased through executive order for the protection of forests and for work relief, supplying jobs to the local communities in reforestation, fire fighting, road building, and soil restoration projects. Loblolly pine became the species of choice, as it was rapid to grow and regenerate.

In more recent years, timber, fire, wildlife, and T&E management have worked well together to restore and maintain fire-associated ecosystems including longleaf pine. The Forest began an active prescribed fire program in 1944 and since 1956, an average of 25,000 acres have been burned annually. Prior to Hurricane Hugo, which destroyed 90% of the Forest canopy, the Forest contained 36,100 acres in longleaf pine (14%) and 112,030 acres in loblolly pine (44%). Today 47,944 acres are in longleaf or mixtures with loblolly pine, which are suitable for restoration. The long-term Forest objective is to increase the longleaf cover type to 53,500 acres, and emphasize longleaf pine ecosystem restoration in Management Area 26, which consists of fairly dry sandy soils supporting flatwoods, savannah, and woodland communities. Future efforts will emphasize first thinnings in stands which were young or which were initiated following Hugo.

184 PHENOLOGICAL PATTERNS OF MATURE LONGLEAF PINE (PINUS PALUSTRIS) TREES ACROSS AN ENVIRONMENTAL GRADIENT

Mary Carol Sheffield (Joseph W. Jones Ecological Research Center, Rt. 2 Box 2324, Newton, GA 31770) Steven B. Jack (Joseph W. Jones Ecological Research Center, Rt. 2 Box 2324, Newton, GA 31770) Daniel J. McConville (Cooperative Forestry Research Unit, University of Maine, Orono, ME (formerly at Jones Research Center))

ABSTRACT: Components of tree crowns and their structural arrangement are important variables in forest ecology studies because they relate to tree growth and productivity through their relationship with light interception. Environmental variables are known to influence crown structure, however the way the environment affects shoot and needle phenology, as well as needle demography, of mature longleaf pine trees is not well understood. This study was initiated to evaluate patterns of shoot and needle phenology in a mature longleaf pine forest both within canopies and across a soil moisture gradient. Towers were used to access mature tree canopies for bi-weekly measurement of shoot and needle elongation, as well as needle abscission, over a growing season at two sites representing the extremes of a natural soil moisture gradient. Soil moisture and predawn leaf water potential were also measured at the two sites. Shoot growth patterns were determinant at both sites. Final shoot length, duration of shoot growth and rate of shoot growth were found to be significantly greater at the wet/mesic site than at the xeric site. Needle initiation, start of linear growth, and growth cessation were also significantly earlier at the wet/mesic site. Needle demography patterns were similar at the sites, and early growing season needle demography (June-September) was related to predawn needle water potential. These findings indicate that water availability influences patterns of phenology in mature longleaf pine trees across the soil moisture gradient, thus productivity at these sites may be influenced by differences in resource availability.

185 HIGHWAY RIGHTS-OF-WAY AS LONGLEAF PINE RESTORATION HABITAT

Philip M. Sheridan (Meadowview Biological Research Station, 8390 Fredricksburg Turnpike; Woodford, VA 22580; Department of Biological Sciences, Old Dominion University, Norfolk, VA 23529-0266) Johnny P. Stowe, Jr. (South Carolina Department of Natural Resources Wildlife Diversity Section; P.O. Box 167; Columbia, SC 29202)

ABSTRACT: Longleaf pine (Pinus palustris Miller) has been extirpated or reduced in extent throughout many parts of its range. We think that highway rights-of-way may offer substantial habitat for longleaf pine planting and could play a significant role in restoration efforts. Desirable attributes of longleaf pine for right-of-way planting include aesthetic appeal and resistance to insects, disease, fire, and wind throw. We highlight a pilot project in Virginia that may be applicable range wide for longleaf pine. Five longleaf pine reintroductions have been done with the Virginia Department of Transportation utilizing cloverleafs and a wetland mitigation site. The plantings have been done as part of an integrated ecosystem restoration and include attractive, indigenous, associate plants. Survival rate of longleaf pine at the first cloverleaf site planted in 1999 are 84% after one growing season. The plantings also represent a 25% increase in the native Virginia longleaf pine population. This project indicates that substantial gains in longleaf pine restoration may be obtained by utilizing appropriate habitat on highway rights-of-way.

INTRODUCTION Longleaf pine historically occurred on about 92 million acres in the southeastern United States (Frost, 1993). Today, less than 3 million acres of this once expansive forest remain. In Virginia, the northern limit for the species, less than 800 acres remain of an estimated original 1.5 million acre longleaf pine forest (Sheridan et al., 1999a; Frost, In Prep.). Given such drastic declines in the distribution of a major forest species, we sought to develop novel approaches to maintaining or increasing acreage for restoration purposes.

This paper highlights a pilot project with the Virginia Department of Transportation (VDOT) to reintroduce longleaf pine within its historic range on appropriate habitat on highway rights-of-way. Highway rights-of- way may offer substantial habitat for longleaf pine planting and could play a significant role in restoration efforts.

MATERIALS AND METHODS Virginia longleaf pine seeds were collected and propagated by Meadowview Biological Research Station (Sheridan et al., 1999b). Longleaf pine seedlings were donated to students at Potomac Elementary School in King George County, Virginia, as part of their rare plant restoration program under the Toyota Tapestry Grant. Students raised their seedlings in 6-inch pots and maintained them at constant moisture through a bottom watering system.

Five sites were selected for longleaf pine restoration. Three of the sites were selected by a VDOT roadside development manager and involved small-scale roadside plantings of Meadowview longleaf stock (n = 10, 25, and 25) in Portsmouth, Sufolk, and Virginia Beach, Virginia, respectively. The 2 other plantings were located in a cloverleaf in Prince George County, Virginia (n = 217), and a wetland mitigation site in Greensville County, Virginia (n = 590).

The Prince George County longleaf pine planting was done in January 1999 with bare root and container stock. Ground cloth and mulch were placed around each longleaf pine seedling to control competition. A bamboo stake with survey tape was also placed close to each seedling so that maintenance crews would avoid mowing the seedlings.

The Greensville County wetland mitigation site received 2 longleaf pine plantings: 500 seedlings planted by Meadowview biologists, and 90 seedlings raised by Potomac Elementary School students. Seedlings were planted in December, 1999 in mineral soil without mulch or groundcover. The seedlings planted by the Potomac Elementary School students were delivered as potted plants to the mitigation site and planted as container stock (6 inch pots) while the Meadowview material was bare root. Native Virginia yellow pitcher plant, Sarracenia flava L., was also planted by Meadowview biologists (adult plants, n = 361) and Potomac

186 Elementary School students (one-year-old, 1/0 seedlings, n = 365) in May of 2000 as part of an integrated ecosystem restoration.

RESULTS Survival of longleaf pine planted at the Prince George County cloverleaf site was 84% after 1 growing season (Sheridan, 2000). Longleaf pine survival at the Greensville County wetland mitigation site was 61% for the Potomac Elementary School seedlings and 54% for the Meadowview seedlings. No survival data was provided for the other VDOT sites. Survival of S. flava at the Greensville County site was 82% for adult plants and 44% for seedlings.

DISCUSSION These longleaf pine plantings on highway rights-of-way in Virginia represent an 11% increase in the native population of 4,432 trees (Sheridan et al., 1999a), and the reintroduction of this species to 2 historic counties (Harris, 1999; Tennant, 2000). The situation for longleaf pine in Virginia was so dire that significant increases in population size were easily achieved by roadside and mitigation site planting. Can such dramatic increases in longleaf pine populations be achieved in other states by planting along highway rights-of-way? We suspect that a careful evaluation of appropriate habitat along highway corridors will disclose significant acreage for longleaf pine planting. The relative success (84% survival) of longleaf pine planting in Prince George County may be attributable to the effective control of competition provided by mulching the seedlings (Haywood, 2000). Longleaf pine survival rates at the Prince George site would have been even higher but for losses resulting from accidental mowing in the cloverleaf. The low survival rate at the Greensville County mitigation site is explained by late site preparation by VDOT, soil subsidence around seedlings, and subsequent high mortality rates. However, we were willing to accept these high mortality rates since we have found very few agencies with protected land in Virginia willing to plant longleaf pine. The VDOT has been a leader in the effort to restore native Virginia longleaf pine.

Many DOT projects have used exotic ornamentals to "beautify" highway rights-of-way. These species often are poorly adapted to local conditions and require expensive, polluting, and time-consuming efforts to maintain; they also may become nuisance, invasive species. In contrast, longleaf pine is a signature tree of the southeastern United States that is well adapted to local conditions and requires little maintenance outside the seedling phase. Our longleaf pine restoration work on highway rights-of-way in Virginia has focused on incorporating a number of rare, native associate species in these plantings as part of regional restoration efforts (Sheridan and Penick, In Press).

One of the primary values of planting longleaf pine on highway rights-of-way is to act as a catalyst to bolster large-scale restoration efforts. Longleaf pine and associate species could be planted along highway rights-of- way and at rest areas and visitor centers. Rest areas and visitor centers provide the appropriate, safe setting for short trails with detailed interpretive signage to satisfy public curiosity and disseminate information.

Although highway rights-of-way fragment wildlife habitat in rural areas, they may act as refugia in urban areas where longleaf pine has been locally extirpated. These refugia along highway rights-of-way provide appropriate habitat for longleaf pine persistence and regeneration through the control of competitors and exposure of mineral soil by mowing operations. Careful consideration must be made in the choice of native associate plant species used in longleaf pine restoration along highway rights-of-way. We are using longleaf pine planting sites in Virginia to help restore a number of other rare plant species (e.g. S. flava). Plantings by wildlife departments in demonstration sites on roadsides have resulted in the formation of ecological traps to bobwhite quail (Colinus virginianus) and other bird species. Higher mortality rates for a threatened bird species have been recorded along rights-of-way than on non-road territories (Mumme et al., 2000). Conservation biologists be consider all options available in the natural palette to produce a result that will not imperil wildlife species.

Involving local school systems in restoring longleaf pine habitat with the VDOT was an important part of our program (Anonymous 2000). Involving the next generation in raising, experimenting with (Sheridan et al., 2000), and reintroducing rare native plants is important to the future of biological conservation. This pilot project in Virginia highlights how those goals can be achieved.

187 ACKNOWLEDGEMENTS We thank the Toyota Corporation for the Toyota Tapestry Grant to Potomac Elementary School for longleaf pine and yellow pitcher plant restoration. The Virginia Department of Transportation Suffolk and Richmond districts were also extremely helpful and we thank them and their staff for support and assistance. Parents, teachers, master gardeners, and most especially the students at Potomac Elementary School were an important part of this project’s success.

LITERATURE CITED Anonymous. 2000. School children join in effort to propagate plants…rare plants get new start in VDOT wetland. Bulletin - a VDOT Employee Newspaper 66: 1. Frost, C.C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. Pp. 17-43. In: Herman, S.H. ed., The Longleaf Pine Ecosystem: ecology, restoration and management. Proceedings of the Tall Timbers Fire Ecology Conference, No. 18. Frost, C.C. In Preparation. Natural history and vegetation of the Albemarle-Pamlico region, Virginia and North Carolina. Doctoral dissertation, Univ. of North Carolina, Chapel Hill. Harris, C. 1999. Rebirth for “lost” longleaf pine. Independent-Messenger, 102: 1 and 5. Haywood, J.D. 2000. Mulch and hexazinone herbicide shorten the time longleaf pine seedlings are in the grass stage and increase height growth. New Forests 19: 279-290. Mumme, R.L., S.J. Schoech, G.E Woolfenden, and J.W. Fitzpatrick. 2000. Life and death in the fast lane: demographic consequences of road mortality in the Florida scrub-jay. Conservation Biology 14: 501- 512. Sheridan. P. 2000. A method for planting longleaf pine, Pinus palustris Miller, on highway rights-of-way. Virginia Journal of Science 51: 129. Sheridan, P., J. Scrivani, N. Penick, and A. Simpson. 1999a. A census of longleaf pine in Virginia. Pp 154- 162. In: Kush, J.S., comp. Longleaf pine: a forward look. Proceedings of the Second Longleaf Alliance Conference; 1998 November 17-19; Charleston, SC. Longleaf Alliance Report No. 4. Sheridan, P., N. Penick, A. Simpson, and P. Watkinson. 1999b. Collection, germination, and propagation of Virginia longleaf pine. Pp 151-153. In: Kush, J.S., comp. Longleaf pine: a forward look. Poceedings of the Second Longleaf Alliance Conference; 1998 November 17-19; Charleston, SC. Longleaf Alliance Report No. 4. Sheridan, P., R. Horman, S. Horman, S. Gilbert, A. Keeton, and M. Schmutte. 2000. Rare plants in the classroom; Potomac Elementary School and the Toyota Tapestry Grant. Virginia Journal of Science 51: 110, 130. Sheridan, P. and N. Penick. In Press. Highway rights-of-way as rare plant restoration habitat in coastal Virginia. In: The Seventh International Symposium on Environmental Concerns in Rights-of-Way Management; 2000 September 9-13; Calgary, Alberta, Canada. Tennant, D. 2000. A Bog’s Life. The Virginian Pilot. E1.

188 LONGLEAF PINE ACTIVITIES IN VIRGINIA: 1998 - 2000

Philip M. Sheridan (Meadowview Biological Research Station, 8390 Fredericksburg Turnpike, Woodford, VA 22580; Blackwater Ecologic Preserve, Department of Biological Sciences, Old Dominion University, Norfolk, VA 23529-0266) Robert A. S. Wright (Parsons Transportation Group, De Leuw. Cather & Company of Virginia, 11320 Random Hills Road, Suite 100, Fairfax, Virginia 22030)

ABSTRACT: Longleaf pine is a very rare tree in Virginia with only 4432 trees remaining in four wild stands in the state. Most of the trees are found on only two sites, Blackwater Ecological Preserve (n=2139) and South Quay (n = 2251). The Blackwater Preserve is a protected area while the South Quay tract is owned by International Paper. In order to protect and manage these remaining resources several avenues of research have been pursued. These efforts include: a reproductive model for the Blackwater Ecological Preserve calculating time to stand regeneration based on stem diameter, growth rate, and seed yield; Virginia Department of Forestry initiation of a native Virginia longleaf orchard; a plan for a functional native Virginia longleaf pine seed production area utilizing the South Quay tract with projected seed yields; our efforts involving elementary school students raising, planting, and doing research on longleaf pine; successful efforts to root longleaf pine from needle fascicles; pilot projects with landowners.

INTRODUCTION Longleaf pine (Pinus palustris Mill.) reaches its northern limit in Virginia and is considered extremely rare in the state (Killeffer 1999). Frost (2000) estimated longleaf pine forests in Virginia occupied as much as 1.5 million acres at the time of European settlement. Today only 4432 trees remain on less than 800 acres (Sheridan et al. 1999b). To reverse these drastic declines in population size a recovery effort has been initiated. This paper highlights longleaf pine restoration and conservation efforts in Virginia from 1998 - 2000.

Reproductive Model The only protected longleaf pine ecosystem in Virginia is the Blackwater Ecological Preserve in Isle of Wight County. Reproduction of longleaf pine at the preserve is currently inadequate to produce grass stage seedlings (Sheridan et al. 1999b). When will the longleaf stand at the Blackwater Ecological Preserve be capable of natural regeneration? If longleaf pine fecundity and growth rate can be calculated then a prediction of when regeneration will occur can be made.

Platt et al. (1988) found a rather close correlation between the stem diameter of longleaf pine and the number of cones produced. Boyer (1998) determined that 360 cones per acre are needed just to get the first successful seedling and Croker (1973) calculated that the average number of seeds per longleaf cone is 50. This translates into a requisite 18,000 seeds/acre for regeneration. Sheridan et al. (1999b) performed a calculation synthesizing these authors results and longleaf pine seed-fall data collected at the preserve and found that only a third the amount of seed necessary for successful regeneration was produced in 1987. This inadequate seed production is caused by the lack of cones produced by small diameter trees. Deficiencies in seed production were still evident in 1998 since only 9 grass stage seedlings were counted in the census (Sheridan et al. 1999b), many possibly planted in a previous study (Plocher 1993). Seed production in 1998 was estimated at only 3000 seeds/acre if one assumes 15 trees/acre from census data and 4 cones per tree based on a 20 cm. diameter tree (Platt et al. 1988). Hence current lack of regeneration at the preserve can be explained through calculations with fecundity data.

Increment core data of longleaf pine at the preserve may be used to calculate average growth rate. Average growth rate can then be used to estimate when trees will have stem diameters large enough for successful regeneration. The average diameter of the stand in the 1998 census was 20 cm. (Sheridan et al. 1998b). Plocher (1993) stated that the entire stand was cut and burned from 1955-1957 and either planted in Louisiana longleaf or naturally regenerated. Assuming 40 years of growth by 1998 and an average stem diameter of 20 cm. the average rate of growth was calculated as 0.5 cm./year. Validation of this assumption was provided by increment core data collected in 2000 that disclosed an average age of 43 years (range 38-46 growth rings, n=12) and growth rate of 0.54 cm/year.

189 Therefore, a reproductive model of time to natural regeneration at the Blackwater Ecological Preserve can be constructed based on the following parameters.

Assumptions - 360 cones/acre required to obtain first seedling (Boyer 1998) - cones produce an average of 50 seeds (Croker 1973) - longleaf pine with stem diameter of 45 cm will produce 24 cones/tree (Platt et al. 1988) - even distribution of trees - no mortality - regular prescribed fire

Data - 2139 longleaf pine cover 143 acres at preserve - = average density of 15 trees/acre - growth rate has averaged 0.54 cm/year - average stem diameter of 20 cm in 1998

Calculation to Time of Regeneration 15 trees/acre x 24 cones/tree = 360 cones/acre 45 cm - 20 cm = 25 cm to grow 25 cm/1 x 1 year/0.54 cm = 46 years

Adjustments to Model This reproductive model assumes an even distribution of longleaf pine on the Blackwater Ecological Preserve. However, longleaf pine tends to be patchy in distribution at the preserve. Higher densities of longleaf pine caused by this patchy distribution may accelerate regeneration estimates. On the other hand aspects of stand structure may also affect local regeneration. Cone production will be higher on dominant and co-dominant trees, which are larger than the 20 cm average. Competition from surrounding longleaf, loblolly, and pond pine may limit cone production on dominant trees. Management efforts may be required to provide a stand structure more conducive to natural regeneration (John Scrivani, pers. comm.).

An additional factor to also consider is the possibility of an exceptional mast crop producing enough seeds for regeneration prior to the calculated regeneration time. However, bumper crops of such magnitude are very infrequent and have only been recorded twice in the past 50 years (Boyer 1998).

Mortality effects, however, may significantly retard estimates of time to regeneration. For example, annual mortality rates of longleaf pine range from 0.08 for trees of 2.54 cm to less than 0.01 for trees greater than 15 cm (Quicke et al. 1997). Quicke et al. (1997) however excluded catastrophic events, such as fire, from their model. Fire is an essential component of restoration and maintenance of the longleaf pine ecosystem and should be considered in such a model. Catastrophes have been included in models of other rare species (Root 1998).

Platt et al. (1988) reported mortality of 1.75-2.65% for longleaf pine in a fire maintained preserve in south Georgia. This can be contrasted with the rather high 10.3% mortality rate of adult longleaf pine at the Blackwater Ecological Preserve after burning (Plocher 1993). These high mortality rates of longleaf pine at Blackwater Ecological Preserve may reflect catastrophic effects on a fire-suppressed system. The preserve is now being burned on a regular basis and a better assessment of mortality can now be made. This assessment may be made by comparison of plot survival to census data. Once mortality data is obtained the model may be adjusted to recalculate time to stand regeneration

Virginia Department of Forestry Initiates Longleaf Orchard The Virginia Department of Forestry is carefully evaluating remaining, native longleaf pine trees in Virginia for selection as part of a seed orchard. Scions from native trees have been taken and grafts have been successfully made. These ramets have been planted in the field at the Department’s New Kent Forestry Center. The goal is to have a final orchard of 50 or more trees supplying high quality, Virginia source, seedlings for local forestry demands (Anonymous 2000a).

190 Functional Native Virginia Longleaf Pine Seed Production Area While the Virginia Department of Forestry is preparing a native longleaf pine seed orchard it may be decades before the facility is in seedling production. In the interim the native longleaf pine site in Suffolk City at South Quay (Sheridan et al., 1999b) could serve as a functional orchard, or seed production area, for longleaf pine seed (Sheridan 1999). Two thousand two hundred and fifty one longleaf pine occur on 459 acres at the South Quay tract. Ninety percent of these trees (n = 2030) occur on only two tracts totaling 275 acres. Calculations of seed production based on stem diameter at these two tracts result in an estimated possible annual yield of 211,950 seeds (Table 1).

Initial cone collections at the South Quay tract in 1997 were inefficient and resulted in a low yield per cone due to an early maturing cone crop in Virginia. Since 1997, cone collections have been performed earlier in the season to determine optimal cone collection time. In late September 2000 mature, green, unopened cones were harvested and an average yield of 60 seeds per cone was obtained (Table 2), closely matching the average value of 50 seeds per cone reported by Croker (1973).

Cones are simply and efficiently collected at South Quay using a 30-foot telescoping forestry pole to knock cones off the tree. This low technology approach avoids damage to the sensitive area through heavy equipment and results in enough seedlings to meet current (non-forestry) demand for longleaf pine restoration in Virginia. A more aggressive harvest of cones could also meet forestry demands. An important point to consider is that the South Quay tract is one of the last native longleaf stands in Virginia currently capable of supplying seed for restoration. As environmental stewards we have a responsibility to ensure that native Virginia longleaf pine is propagated and planted in the state in preference to other genotypes.

Educational Efforts Potomac Elementary School, in King George County, Virginia was awarded both the Toyota Tapestry and Dominion Virginia Power Partnership Grant to help restore Virginia's longleaf pine and pitcher plants. Students studied germination rates of longleaf pine and reintroduced their 95 seedlings to a county from which longleaf pine had been extirpated in Virginia. The reintroduction occurred at a Virginia Department of Transportation wetland mitigation site (Harris 1999, Anonymous 2000b, Sheridan et al. 2000a, Tenant 2000) and represented a significant recovery effort for longleaf pine in the state. Propagation and research efforts continue in school yard "mini-nurseries".

Regenerating Longleaf Pine from Needle Fascicles Our goal is to capture the entire native Virginia longleaf pine genome on one preserve. Since non-native longleaf pine trees have been co-planted with indigenous material in some stands truly native provenance may be compromised through seed harvests. If needle fascicles from mature trees can be rooted and regenerated then an entirely native longleaf pine genome can be established on a preserve. Fascicles from grass stage longleaf pine up to three years old can be successfully rooted within one month. The technique involves placing the fascicles upright in a 1cm deep solution of 100 ppm IBA followed by maintenance in a nutrient solution of 60 ppm Boric Acid, 40 ppm Ammonium Nitrate, and 20 ppm Thiamine-HCl (Nelson et al. 1992, Sheridan et al. 2000b). Rooting rates with Virginia longleaf pine have been as high as 65% (Table 3). Only 5% of rooted fascicles have entered the shoot initiation phase while many show evidence of apical meristem development after 10 months. Survival and growth of rooted fascicles, called "needlings", is comparable with conventional seedlings (Lott and Nelson 1999). Although scientists have not been able to get fascicles from longleaf pine over 3 years old to root, researchers eventually hope to find a successful technique to make fascicles from older longleaf pine trees root and regenerate.

Pilot Projects with Landowners Test sites are important in convincing landowners that longleaf pine is a potential tree for reforestation. Experimental plantings have been done to show landowners that longleaf pine can be successfully planted in Virginia and that growth is satisfactory.

LITERATURE CITED Anonymous. 2000a. Preserving the native longleaf seed source in eastern Virginia. Virginia Forests Magazine 55: 14-20

191 Anonymous. 2000b. School children join in effort to propagate plants...rare plants get new start in VDOT wetland. Bulletin, a VDOT employee newspaper 66: 1. Boyer, W.D. 1998. Longleaf pine regeneration and management: an overstory review. p. 14-19. In. Kush, J.S., comp. Ecological Restoration and Regional Conservation Strategies. Proc. of the Longleaf Pine Ecosystem Restoration Symposium, Pres. at Soc. for Ecological Restoration Ninth Annual International Conference. Longleaf Alliance Report No. 3. Auburn University, AL. Croker, T.C., Jr. 1973. Longleaf pine cone production in relation to site index, stand age, and stand density. Res. Note SO-156. USDA Forest Service, Southern Forest Experiment Station, New Orleans, LA. 3 pp. Frost, C.C. 2000. Natural history and vegetation of the Albemarle-Pamlico region, Virginia and North Carolina. Ph.D. diss., Univ. of North Carolina, Chapel Hill. Harris, C. 1999. Rebirth for "lost" longleaf pine. The Independent Messenger, Dec. 26: 1, 5. Killeffer, S.E. 1999. Natural Heritage Resources of Virginia: Rare Vascular Plants. Natural Heritage Technical Report 99-11. Virginia Department of Conservation and Recreation, Division of Natural Heritage, Richmond, Virginia. Unpublished report. April 1999 36 pages plus appendices. Lott, L.H. and C.D. Nelson. 1999. Longleaf needle-derived rooted seedlings after 5 years in the field. Proceedings of the 25th southern forest tree improvement conference. New Orleans, LA; 11-14 July 1999. P. 219 Nelson, C.D., Z. Linghai, J.M. Hamaker. 1992. Propagation of loblolly, slash and longleaf pine from needle fascicles. Tree Planters Notes 43: 67-71. Platt, W.J., G.W. Evans, and S.L. Rathbun. 1988. The population dynamics of a long-lived conifer (Pinus palustris). The American Naturalist 131(4):491-525. Plocher, A.E. 1993. Population dynamics in response to fire in Quercus laevis - Pinus palustris barrens and related communities in southeast Virginia. Ph.D. dissertation. Old Dominion University, Norfolk, VA. Quicke, H.E., R.S. Meldahl, and J.S. Kush. 1997. A survival rate model for naturally regenerated longleaf pine. Southern Journal of Applied Forestry 21(2):97-101. Root, K. 1998. Evaluating the effects of habitat quality, connectivity, and catastrophes on a threatened species. Ecol. Appl. 8:854-865. Sheridan, P., N. Penick, and A. Simpson, P. Watkinson. 1999a. Collection, germination, and propagation of Virginia longleaf pine. Pp. 151-153. In: Kush, J.S. comp. Longleaf Pine: A Forward Look. Proc. Second Longleaf Alliance Conference. Longleaf Alliance Report No. 4. Auburn University, AL. Sheridan, P., J. Scrivani, N. Penick, and A. Simpson. 1999b. A census of longleaf pine in Virginia. Pp. 154- 162. In: Kush, J.S. comp. Longleaf Pine: A Forward Look. Proc. Second Longleaf Alliance Conference. Longleaf Alliance Report No. 4. Auburn University, AL. Sheridan P., R. Horman, S. Horman, S. Gilbert, A. Keeton, and M. Schmutte. 2000a. Rare plants in the classroom; Potomac Elementary School and the Toyota Tapestry grant. Virginia Journal of Science 51: 130. Sheridan, P., K. Nesius, and L. Everett. 2000b. Rooting longleaf pine, Pinus palustris Miller, from needle fascicles. Virginia Journal of Science 51: 99. Tennant, D. 2000. A bog's life. The Virginian-Pilot, May 16: E1, E4.

ACKNOWLEDGEMENTS Special thanks to Toyota Corporation and Dominion Virginia Power for grants awarded to Potomac Elementary School. Teachers, parents, mentors, master gardeners, and especially the students did an outstanding job in their work restoring longleaf pine. A grant from the Virginia Academy of Science through the Small Project Research Award supported the project to root longleaf pine from needle fascicles. Additional thanks to Robert Wright for research assistance and John Scrivani from the Virginia Department of Forestry for review and comments.

192 Table 1. Calculated cone and seed yield at South Quay longleaf pine stand in Suffolk City, Virginia. ______Tract 1 Tract 2 DBH (cm) # Trees # Cones/tree Total # Trees # Cones/tree Total ______

3 8 0 0 54 0 0 5 10 0 0 17 1 0 8 20 0 0 219 0 0 10 27 1 27 232 1 232 13 22 1 22 168 1 168 15 23 1 23 170 1 170 18 12 1 12 159 1 159 20 26 2 52 144 2 288 23 27 4 108 130 4 520 25 28 4 112 101 4 404 28 22 4 88 75 4 300 30 19 6 114 42 6 252 33 10 6 60 26 6 156 36 12 6 72 22 6 132 38 8 12 96 13 12 156 41 9 12 108 7 12 84 43 2 18 36 4 18 72 46 5 18 90 48 3 18 54 51 1 24 24 53 2 24 48

Total 296 1146 1737 3093

Calculation Tract Total cones x Avg. seed/cone Total seeds

1 1146 50 57,300 2 3093 50 154,650 Total 211,950 ______

Table 2. Longleaf pine seed yield per cone and date of harvest at South Quay, Suffolk City, Virginia. ______

Date # Cones # Seeds Avg. seed/cone ______

11/3-11/14/97 121 2614 22 10/15/99 51 1293 25 9/29/2000 47 2855 61 ______

193 Table 3. Percentage of fascicles from two year old longleaf pine seedlings that developed roots. ______

Seedling # # Fascicles % rooted Comments ______

1 34 0 ------2 57 0 ------3 31 65 callus on unrooted fascicles 4 16 0 callus on unrooted fascicles 5 29 24 callus on unrooted fascicles 6 41 10 mix of green and dead fascicles remaining 7 40 0 all dead 8 52 3 50% fascicles still green 9 40 20 50% fascicles still green 10 27 44 44% fascicles still green ______

194 INSECT POLLINATORS OF RARE PLANTS IN APALACHICOLA LONGLEAF PINE FOREST

Theresa L. Pitts-Singer (USDA Forest Service, Forestry Sciences Laboratory, 320 Green St., Athens, GA 30602) James L. Hanula (USDA Forest Service, Forestry Sciences Laboratory, 320 Green St. Athens, GA 30602) Joan L. Walker (USDA Forest Service, Department of Forestry Resources, Clemson University, Clemson, SC 29634)

ABSTRACT: As a result of human intervention, longleaf pine forests in the Southern United States have been lost or drastically altered. Many of the plants and animals that historically occupied these forests persist only in forest remnants and have been placed on the Endangered Species List. In order to safeguard endangered or threatened species in such forests, we need to better understand their natural history and ecology. We have performed studies of the pollinators of three rare plant species that occur in the longleaf pine habitat of the Apalachicola National Forest. These plants are Harperocallis flava (F. Amaryllidaceae), Macbridea alba (F. Lamiaceae) and Scutellaria floridana (F. Lamiaceae). Although much emphasis has been placed on the response of plants to forest management practices, there is little information concerning the impact of forest management on the survival and availability of insect pollinators of these rare plants. We have identified insects that visit the three plant species and evaluated their importance as pollinators. Eventually, we hope to determine if pollinators are limited, if low numbers of flowers fail to attract or maintain pollinators, and if there is competition for pollination service between rare flowers and other more common local ones. The results of this study can be used to make I inferences about how forest management practices impact the survival of rare plants as well as their pollinators. _____

Managers of public lands are obligated to use management practices that preserve threatened and endangered species (Endangered Species Act of 1973). Previous studies of potential management effects on rare plants have considered the response of the plants to management practices (e.g., Hessl & Spackman 1995; Brewer 1999; Lesica 1999; and references therein), but have not evaluated effects on the pollinators. We are interested in the pollinator-plant relationships of Harper's beauty (Harperocallis flava; Amaryllidaceae), White birds-in-a-nest (Macbridea alba; Lamiaceae), and Florida skullcap (Scutellaria floridana; Lamiaceae), which are endemic to the longleaf pine ecosystem in the Apalachicola lowlands of Florida (Kral 1983; Walker 1993) and are federally listed as threatened or endangered.

We observed and recorded insect and spider visitors on the three plant species during the summers of 1999 and 2000. Researcher observations and video-recordings made over one to several days revealed few visitors to flowers. However, for each flower species, one or more of the insect visitors displayed behavior on the flowers that indicated its role as a potential pollinator.

All three plants that we studied on the Apalachicola National Forest are pollinated by bees: bumblebees, megachilids, and halictids. However, flower visitations seemed quite infrequent. Whether the rarity of these plants is affected by low pollinator visitation frequency has not yet been determined experimentally.

Allozyme studies have indicated that H. flava individuals and populations are genetically uniform (Godt et al., 1997). A preliminary study revealed that few insects visited these flowers and that they are self- compatible (L. Wagner & T. Spira, Clemson Univ., SC). Interestingly, we observed several halictid bees collecting pollen from H. flava flowers, especially in 2000. The behavioral activity of halictids on H. flava flowers could have brought collected pollen in contact with the stigma. For a hermaphroditic species like H. flava, visitation by solitary bees may improve out-crossing, or the activity of bees on anthers may dislodge pollen for better wind dispersal (Cane et al., 1992). Nonetheless, with the low genetic variance in H. flava, out-crossing may no longer be very important. On the other hand, the availability of this pollen may be a vital resource for the survival of the solitary bees that harvest it.

Bumble bees were the only insects observed whose size and behavior were suitable for pollinating M. alba. Madsen (1999) found that when insects were excluded from these flowers, almost no seeds were produced. Observations of M. alba flowers in the South Carolina Botanical Garden revealed that bumblebees were very frequent visitors and efficient pollinators (J. L. Walker, personal observation).

195 Our study provides the first information concerning the pollination ecology of S. floridana. We saw that the pollen-collecting behavior of megachilid bees, halictid bees, and one bumble bee were appropriate for pollination of this flower species. Carpenter bees were seen to rob nectar by piercing the base of the corolla. Insect exclusion experiments and comparative studies with closely related common skullcaps are necessary for further understanding of the pollination requirements of this species.

Each of the plants that we have studied has flowering periods that occur in response to fire. This flowering response suggests that fire timing is important to ensure that flowering occurs when pollinators are available. Thus, the life cycle of important pollinators must also be made known. Furthermore, we need information concerning other important food sources and nesting materials that are required to maintain pollinators in the area. Knowledge of insect pollinators and of the interdependence of insects and these protected plant species is essential for developing management strategies that ensure the viability of both the plants and insects.

We extend our gratitude to the personnel of the Apalachicola National Forest Ranger Station for their support of this project, especially Louise Kirn. David Jenkins of the University of Georgia was a valuable field assistant in summer 1999, and his contributions are much appreciated.

REFERENCES Brewer, J.S. 1999. Effects of competition, litter, and disturbance on an annual carnivorous plant (Utricularia juncea). Plant Ecol. 140: 159-165. Cane, J.H., Buchmann, S.L. & LaBerge, W.E. 1992. The solitary bee Melissodes thelypodii thelypodii Cockerell (Hymenoptera: Anthophoridae) collects pollen from wind-pollinated Amaranthus palmeri Watson. Pan-Pacific Entomol. 68: 97-99. Godt, M.J.W., Walker, J. & Hamrick, J.L. 1997. Genetic diversity in the endangered lily Harperocallis flava and a close relative, Tofieldia racemosa. Conserv. Biol. 11: 361-366. Hessl, A. & Spackman, S. 1995. Effects of fire on threatened and endangered plants: an annotated bibliography. Information and Technology Report 2. U.S. Department of the Interior, Washington, D.C. 55 pp. Kral, R. 1983. A report on some rare, threatened, or endangered forest-related vascular plants of the south. UDSA Forest Service Tech. Pub. R8-TP, Vols. 1 & 2. Lesica, P. 1999. Effects of fire on the demography of the endangered, geophytic herb Silene spaldingii (Caryophyllaceae). Amer. J. Bot. 86(7): 996-1002. Madsen, D.L. 1999. Seed production and germination studies of Macbridea alba. Master's Thesis, Clemson University. Walker, J. 1993. Rare and vascular plant taxa associated with the longleaf pine ecosystems: patterns in and ecology. Proc. Tall Timbers Fire Ecol. Conf. #18, pp. 105-125.

196 ABOVE- AND BELOW-GROUND GROWTH OF LONGLEAF PINE IN RESPONSE TO THREE PRESCRIBED BURNING REGIMES

Mary Anne Sword (U.S.D.A. Forest Service, Southern Research Station, Pineville, LA) Eric Kuehler (U.S.D.A. Forest Service, Southern Research Station, Pineville, LA)

ABSTRACT: Maintenance of longleaf pine ecosystems requires repeated fire. Past research has indicated that in some situations, regular burning decreases longleaf pine productivity. Growth reductions may be attributed to fire-induced loss of leaf area. It is possible that the loss of leaf area is a function of both fire intensity and the stage of flush development at the time of burning. The objective of this study is to evaluate the effect of prescribed burning at different stages of branch development on the periodic growth, leaf area dynamics, root elongation and root carbohydrate relations of 45-year-old longleaf pine. We hypothesize that the potential for fire damage to new fascicles will be reflected in leaf area responses to seasonal burning, and that periodic growth, root elongation and root carbohydrate relations will be related to leaf area dynamics.

The study is being conducted on the Palustris Experimental Forest, Rapides Parish, LA, in an existing long- term experiment where seasonal burning treatments have been applied biennially since establishment. Since the long-term study is unevenly stocked, the present study is being conducted in two replications of tree clusters that are of equivalent stocking. Prescribed burning treatments were applied in March, May and July 1998. Tree growth, leaf area index, root elongation and root carbohydrate concentrations were measured periodically during March 1998 through June 2000. Seasonal patterns of leaf area, tree growth, root elongation and root carbohydrate concentrations will be presented. Relationships between season of burn, leaf area dynamics and growth processes in tree clusters will be discussed.

INTRODUCTION Controversy exists concerning the sustainability of southern pine forests that experience repeated prescribed fire. Although some research indicates that routine prescribed fire does not negatively affect southern pine productivity, other studies show that regular burning reduces tree growth (Boyer 1987, Brockway and Lewis 1997). As a result, long-term use of repeated prescribed fire requires understanding of its interaction with forest productivity.

Carbon fixation and allocation in the crown of trees treated with prescribed fire are affected by fire intensity. For example, crown scorch and premature foliage senescence have been associated with reductions in both diameter growth and root carbohydrate concentrations for two years after prescribed fire (Johansen and Wade 1987, Sword and Haywood 1999).

Growth reductions have also been observed with negligible foliage loss in response to repeated burning (Johansen and Wade 1987, Boyer 1987). The impact of minor foliar damage on tree growth may be a function of two crown variables. First, the stage of fascicle development at the time of prescribed fire and the sensitivity of new foliage to heat may influence fire damage to current-year foliage. Second, the stage of seasonal leaf area dynamics at the time of prescribed burning may influence the risk of fire-induced damage to stand leaf area. Since carbon allocation to sinks such as stem and root diameter growth, new root initiation and starch storage occurs seasonally, damage to the foliar source of carbon for these sinks may limit tree growth.

We hypothesize that prescribed fire damages foliage and subsequently, reduces leaf area which decreases carbon fixation in the crown and carbon allocation to stem and root growth. Furthermore, this effect depends on fascicle development and leaf area at the time of prescribed burning, and the seasonal dynamics of carbon allocation to stem and root growth. Our objective is to present the seasonal dynamics of leaf area and root carbohydrate concentrations in response to three prescribed burning regimes, and evaluate interaction between leaf area and root carbohydrate concentrations in a 45-year-old longleaf pine stand. As the study continues, similar relationships between leaf area and other important carbon sinks will be evaluated.

MATERIALS AND METHODS The study is being conducted in a long-term experiment on the Palustris Experimental Forest, Rapides Parish, Louisiana. The study site was regenerated with longleaf pine by the seed tree silvicultural system between

197 1950 and 1956. Prescribed fire was applied biennially between 1966 and the present. The soil is a complex of three well-drained, very fine or fine sandy loam soil series. The original study consisted of four treatments assigned to sixteen 36.6 x 36.6 m2 plots in a randomized complete block design with four blocks. Blocking was based on topography. Treatments were biennial prescribed burning in winter (March), spring (May) or summer (July).

In the absence of prescribed fire for 35 years, longleaf pine did not persist on the non-burned plots. Also, the burned plots are unevenly stocked with longleaf pine. Thus, the experimental design of the present study was modified from that of the original study. Specifically, measurements were conducted in 3- to 4- and 6- to 7- tree clusters in two replications of three prescribed burning treatments using a completely random design. For comparison to non-burned trees, measurements were conducted in similar tree clusters in one non-treated plot of an adjacent long-term study. Tree age, stocking and site conditions on the non-burned plot were similar to those in the present study.

Leaf area index (LAI) was monitored monthly with two LiCor LAI-2000 Plant Canopy Analyzers (LiCor Inc., Lincoln, Nebraska). At each measurement interval, 10 readings, equidistant and permanently marked along a transect, were taken in each 6- to 7-tree cluster. Above-canopy readings were taken in an adjacent field. Above- and below-canopy readings were used to determine cluster LAI. Root carbohydrate relations were measured monthly in 3- to 4-tree clusters throughout the two-year burning cycle. Five soil cores were extracted from random locations in the periphery of 3- to 4-tree clusters. Cores were pooled and roots (>2>5 mm in diameter) were removed, washed, lyophilized and ground with a Wiley mill (1 mm2 mesh). Concentrations of root starch, sucrose and glucose were determined enzymatically by the method of Jones et al. (1977) with modifications for loblolly pine.

RESULTS AND DISCUSSION Distinct patterns of leaf area were observed annually (Figure 1). However, leaf areas did not differ by season of prescribed burn. Maximum LAI occurred in September 1998 and 1999, and minimum LAI occurred in February 1998 and January 1999. Differences between maximum and minimum unadjusted LAI in each large tree cluster were calculated for 1998 and 1999. Values for 1998 and 1999 represent leaf area growth in 1997 and 1998, respectively. The leaf area of foliage produced in 1997 was greater than that produced in 1998.

LAI 3

Maximum in September Maximum in September 2.5 Minimum in February Minimum in January

2

1.5

1 SPR98 SUM98 FAL98 WIN99 SPR99 SUM99 FAL99 WIN00 SPR00

Figure 1. Leaf area index (unadjusted) of 45-year-old longleaf pine in three prescribed burning regimes.

Root carbohydrate concentrations were not affected by season of prescribed burn but did vary seasonally (Figure 2). In early spring when environmental limitations were low, significant relationships between leaf area growth and root carbohydrate concentrations were found. Specifically, correlation coefficients (r) between annual leaf area growth and root starch, glucose and sucrose concentrations in April were 0.5036 (P=0.0664), -0.5242 (P=0.0543) and 0.7811 (P=0.0010), respectively. Thus, the potential exists for silvicultural treatments that negatively affect leaf area to reduce root carbohydrate concentrations, and subsequently root metabolism. As the growing season progressed, however, leaf area growth and root carbohydrate concentrations were not correlated. Both 1998 and 1999 were characterized by water deficits and high temperature during the growing season. Limiting environmental conditions

198 may have directly affected relationships between leaf area growth and root sink activity during the height of the growing season, or indirectly affected these relationships by reducing leaf area production and therefore, whole-crown carbon fixation in 1998.

30 20 10

0 100 75 50 25 0

400 300 200 100 0 3 4 5 6 7 8 9 10 11 12 1 2 3 4 5 6 7 8 9 10 11 12 1998 1999 ( Non-burned Winter burn Spring burn Summer burn ) Figure 2. Seasonal root carbohydrate concentrations of 45-year-old longleaf pine in three prescribed burning regimes.

As data analyses continue, we will determine critical times of the year when damage to new foliage or canopy leaf area may be detrimental to stem and root growth. With this information, further research will be designed to evaluate the effect of season of prescribed fire on carbon fixation and allocation in younger and more evenly stocked stands.

LITERATURE CITED Boyer, W.D. 1987. Southern Journal of Applied Forestry. 11: 154-157. Brockway, D.G.; Lewis, C.E. 1997. Forest Ecology and Management. 96: 167-183. Johansen, R.W.; Wade, D.D. 1987. Southern Journal of Applied Forestry. 11: 180-184. Jones, M.G.K.; Outlaw, W.H.; Lowry, O.L. 1977. Plant Physiology. 60: 379-383. Sword, M.A.; Haywood, J.D. 1999. Gen. Tech. Rep. SRS-30. USDA For. Serv., SRS: 223-227.

199 EFFECTS OF FIRE INTENSITY ON SHRUB RESPROUTING IN PINE SAVANNA

J. M. Thaxton (Department of Biological Sciences, 508 Life Sciences Bldg, Louisiana State University, Baton Rouge, LA 70803; 225-928-3099) W. J. Platt (Louisiana State University, Baton Rouge, LA 70803)

ABSTRACT: We manipulated pre-fire fuel loads in an upland longleaf pine (Pinus palustris) savanna to test the hypothesis that local variation in fire intensity affects post-fire resprouting patterns of shrubs. Transects were established across elevational gradients from upper slopes to ravines along intermittent streams in a second growth savanna containing both native and exotic shrub species. Shrub species composition and ramet density were determined in randomly selected 1m2 plots within the transects both before and 2 months after early growing season fires in spring 1999. Prior to the fires, each plot was randomly assigned one of four fuel load treatments. Treatments consisted of fine fuel removal, pine needle addition, wood addition and control. During fires, maximum fire temperatures were recorded in each treatment plot. Maximum fire temperatures were significantly higher in pine needle and wood addition plots than in control or fuel removal plots. For all species pooled, post-fire stem density was significantly lower in needle addition and wood addition than in control plots, while fuel removal plots did not differ significantly from controls. Post-fire stem density was reduced below that of pre-fire levels only for wood addition plots. Patterns of resprouting among individual species appeared to be related to the morphology of their underground resprouting organs, with root crown species being more negatively affected than rhizomatous by increased fire intensity. Patterns of response to variation in fire intensity were similar among native and exotic root crown resprouters. These results suggest that variation in fire intensity may affect shrub community dynamics and that the response of species may be predictable as a function of the morphology of their underground organs.

200 USDA FOREST SERVICE STATE & PRIVATE FORESTRY PROGRAMS EFFECTING LONGLEAF PINE ECOSYSTEMS

Don Tomczak (USDA Forest Service, Southern Region, 1720 Peachtree Road, NW, Room 850, Atlanta, GA 30367)

ABSTRACT: The State and Private Forestry program areas of the Forest Service promote the protection and management of forest resources on state, private and community lands, both rural and urban through a collaborative effort with a wide variety of partners.

State and Private Forestry contributes to a regional restoration/conservation strategy for longleaf pine ecosystems through several voluntary, non-regulatory programs. Specific program components effecting longleaf pine include these:

• Forest Stewardship Program • Conservation Reserve Program • Forest Legacy Program • Forest Taxation • Nursery and Tree Improvement • Forest Health • Economic Action Programs • Urban and Community Forestry Programs • Fire Management

The following sections briefly describe each of these programs and their connection to longleaf pine ecosystems. More detailed information can be found at the following website: http://www.southernregion.fs.fed.us/spf/

Forest Stewardship Program The Forest Stewardship Program seeks to encourage long-term stewardship of non-industrial private forestlands. The keystone of the program is the landowner forest stewardship plan, prepared by a professional resource manager and approved by State Forestry agency personnel. The program encourages longleaf ecosystem management when it is compatible with landowner objectives. The Forest Stewardship Program has also been a financial supporter of the Longleaf Alliance.

Program aspirations in regard to longleaf pine: ¾ To ensure landowners are well informed on management of the species. ¾ To build expertise among service providers (foresters, et al) so as to better assist individual landowners.

CONSERVATION RESERVE PROGRAM The Conservation Reserve Program is an environmental improvement program that includes financial incentives to landowners for planting and maintaining trees on designated non-forested acreages. USDA Farm Services Agency administers this program; the Forest Service assists in the forestry aspects. The program attributes superior wildlife benefits to longleaf pine ecosystems in the South, and favors landowners desiring to plant longleaf pine seedlings. As a result, longleaf pine has become the featured tree species in the program.

Program aspirations in regard to longleaf pine: ¾ To encourage landowners to continue to plant longleaf pine on sites suited to the species. ¾ To ensure that there is an adequate seedling supply to meet the needs of the program. (See Nursery and Tree Improvement section below).

201 NURSERY AND TREE IMPROVEMENT State and Private Forestry in Atlanta provides technical assistance and administers funds to State forestry agencies to support production of high-quality and genetically improved seed and planting stock for reforestation purposes. And the National Tree Seed Laboratory in Macon has international expertise in seed collection, testing, conditioning and storing.

Program aspirations in regard to longleaf pine: ¾ To assist in solving the longleaf pine seed shortage problem. ¾ To establish a wider network of seed production areas so there an adequate supply of seed in all transfer zones. ¾ To promote nursery accreditation and seed and seedling certification. ¾ To maximize seedling supply. ¾ To foster increased production from private nurseries of high-quality longleaf seedlings of known provenance.

FOREST LEGACY PROGRAM The Forest Legacy Program was authorized in 1990 to protect private forests from conversion to non-forest uses by means of permanent easements or through direct purchase of forestland. The program allows individual states to decide whether or not to join. Private landowners can voluntarily agree to place conservation easements on their forestlands. States, land trusts, private forest owners, and communities form a partnership with the Forest Service in this effort to sustain a productive forestland base. The program’s popularity is growing.

Program aspirations in regard to longleaf pine: ¾ To get easements on longleaf pine acreages, where designated as a program priority.

FOREST TAXATION Proper tax planning is as important to the private landowner growing trees for profit as the silvicultural techniques he or she uses to grow a profitable timber crop. Unfortunately, forest tax laws are not well understood by most foresters and tax accountants, and even less so by forest landowners. Forestry investments and businesses must have a profit motive for forest management expenses to be deductible from federal income taxes. Congress has provided some favorable advantages and elections to stimulate increased productivity from the nation’s privately owned lands. However, landowners growing forests for non-profit objectives are presently unable to take advantage of these special federal tax provisions. The National Association of State Foresters and other special interest groups have several initiatives to improve the forest tax climate. But changing the federal tax laws is a very difficult process.

Program aspirations in regard to longleaf pine: ¾ To assist in bringing about forest tax laws which are simpler and more favorable to a wider variety of private forest landowners.

FOREST HEALTH PROTECTION Forest Health Protection addresses forest health problems on all forested lands through collaborative partnerships with state, private, National Forest Systems and other federal cooperators.

Program aspirations in regard to longleaf pine: ¾ Although longleaf pine is relatively resistant to pests and diseases, the staff of Forest Health Protection would like to ensure that longleaf pine seedlings are not planted “off-site” and that landowners and service providers are mindful of genetic source of seedlings.

RURAL COMMUNITY ASSISTANCE PROGRAMS The objective of S&PF Rural Community Assistance programs is to enhance the capacity of rural communities to capitalize on opportunities for producing new jobs and income through the use of local timber, wildlife, water, scenic beauty, and other forest resources. Some communities have developed projects using longleaf pine (e.g., pine straw, nursery production, tourism).

202 Program aspirations in regard to longleaf pine: ¾ To help communities utilize the unique properties of longleaf pine ecosystems to economically and socially improve themselves. URBAN AND COMMUNITY FORESTRY PROGRAM The Urban and Community Forestry Program offers technical and financial assistance, and cost share grants, through State forestry agencies to State and regional urban and community forestry councils, local units of government and citizen nonprofit organizations. The purposes of the program are to increase public awareness of the environmental, economic, social and psychological values of urban trees; to improve education and technical support for proper tree planting, maintenance and protection; and to enable the development of self-sufficient local urban and community forestry programs. There are no specific program elements targeted to longleaf pine. Nevertheless, the program affords an avenue for reaching the urban audience with information about longleaf pine ecosystems.

Program aspirations in regard to longleaf pine: ¾ To raise the awareness of the urban populations of the South as to the environmental, economic, social and historical values of longleaf pine ecosystems.

FIRE MANAGEMENT The S&PF Fire and Aviation Management Unit has a major program component in which it provides oversight to the region’s national forests in the use of prescribed burning. Prescribed burning on national forests has steadily increased in recent years, even though forests have more regulatory constraints. Between fiscal years1987 and 1996, national forests prescribe burned an average of 530,000 acres/year (for all ecosystems—not just longleaf). Acreages increased to 706,000 in 1997; 758,000 in 1998; 853,000 in 1999; and then decreased to approximately 750,000 in 2000 (constrained by budget). Prescribed burning targets for 2001 through 2005 are between 925,000 and 1,100,000 acres/year. National forests use fire aggressively to enhance longleaf pine ecosystems on national forests. They seek to use fire across a spectrum of age classes and management regimes in longleaf ecosystems, as well as to incorporate a variety of seasonality of burns into their programs.

Program aspirations in regard to longleaf pine: ¾ To fully incorporate fire into the longleaf pine ecosystems on national forests in the South.

203 OLD-GROWTH LONGLEAF PINE FORESTS – FILLING IN THE BLANKS

J. Morgan Varner, III (Interdisciplinary Ecology Program, Box 118526, University of Florida, Gainesville, FL 32611-8526) John S. Kush (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849)

ABSTRACT: Old-growth forests have been a topic of great interest to scientists, conservationists, and the public, and old-growth longleaf pine forests are no exception. Old-growth forests have been important as habitat for threatened plant and animal species, as benchmarks for natural area management, as clues to past climatic events, as well as aesthetic and cultural resources. While many imperiled forest types have been inventoried, old-growth longleaf pine, until recently, has been unknown. To fill this void, we searched the literature, interviewed regional experts, queried email discussion lists and newsletters, and performed field research to compile a revised list of old-growth longleaf pine acreage and status. The revised list contains 13 stands, ca. 5552 acres, in 5 states. No old-growth stands are known to remain in Louisiana, South Carolina, Texas, or Virginia. Since a similar 1996 report, the acreage declined 43 %, with three stands omitted and two new areas added. Stand ownerships were diverse, with large holdings (> 1000 acres) in both public (Eglin AFB) and private (south Georgia) ownerships. A significant percentage of stands (at least 4 of the 13 stands, ca. 2000 acres) face serious management threats, due to erratic fire management histories. This survey is preliminary; hopes are that Conference-goers will provide missing data and information on this important biological and cultural resource.

INTRODUCTION Old-growth – the word(s) are controversial, both in the myriad of scientific definitions and by the status it represents to conservationists, foresters, and policy-makers. Many definitions of old-growth have been proposed and argued in the literature (see Hunter and White 1995 and Landers and Boyer 1999 for a lengthy discussion). Old longleaf pine forests, nonetheless, do have unique characteristics absent from even mature forests. These include many large trees (> 20” DBH) with only a few primary branches, many large snags, persistent coarse woody debris, high vertical structural diversity, significant percentage of woody biomass in heartwood, living trees with decayed heartwood, and many others (Engstrom and Sanders 1997, Landers and Boyer 1999). These characteristics provide researchers and managers with benchmarks for natural area management, as clues to past climate and disturbance events, as well as valued aesthetic and cultural resources. For native flora and fauna these characteristics provide among the highest quality habitat in existence. To the degree researchers and managers understand these characteristics, we are limited in our management and conservation of all pinelands landscapes.

While many imperiled forest types have been inventoried, old-growth longleaf pine forest acreage, until recently, has been unknown. Means’ (1996) survey of managers, scientists, and conservationists turned up 14 stands, covering ca. 9975 acres. His work, while preliminary, was the first to ask the question “Just how much old-growth remains?” His startling findings of only 14 stands and the paltry 9775 acres told the sad longleaf pine story well. Of the once 92 million acres of longleaf pine forests and savannas that graced the Southeast, approximately 97 percent was lost, and 99.99 percent was no longer old-growth. These facts helped solidify efforts in both conservation and investigation of the extant old-growth longleaf pine stands.

Beginning in 1996-7, we began an informal search of the stands listed by Means. To our surprise and dismay, several sites that remained when Means produced his list were either in serious management predicaments, not old, or had been harvested. Our previous work in Alabama tells this story well (Kush et al. 1999a). To address these shortcomings and the unfortunate situation of old-growth harvesting and management dilemmas, this project was initiated in 1997.

METHODS Our research was focused in three areas: personal contacts, library research, and field visitation. We contacted forest managers, conservationists, landowners, and scientists via telephone, email, the Longleaf Discussion Listserv, the Longleaf Alliance Newsletter, and in person at various conferences and symposia throughout the longleaf pine and old-growth network. Literature searches for citations regarding old-growth

204 longleaf pine forests were performed. Lastly, we visited sites reported to contain old or old-growth longleaf pine stands.

Following these contacts, we listed the following important characteristics of each stand: • Location & Ownership • Approximate size • Management Regime (burning regime, harvesting, etc…) • Past & Current Research • Threats (Non-native species, wildfire, etc…) • Vegetation Type (Sandhill, flatwoods, etc…)

RESULTS & DISCUSSION It don’t look real good… We estimate that only 4637 ac of old-growth longleaf pine forest remains. This represents only 0.002 % of extant longleaf pine forest acreage, and 0.00005 % of the presettlement longleaf pine forest! Our findings show a 53 % decline in old-growth longleaf pine forest acreage from Means’ 1996 estimate. This decline was caused by harvesting and new data. One stand, the 250 ac “Red Hills” or “Silver Creek” Tract in SW Alabama was harvested. Another, the Flomaton Natural Area (#1), had 7 acres killed by wildfire and another 2 harvested to widen a US Highway. Eglin AFB’s acreage has been reduced by fire damage, and newer estimates by the DoD place its figure lower (McWhite et al. 1999). Similarly, both the Boyd Tract (NC) and the Big Pine Tract (FL) have been degraded due to severe fire damage. After reviewing literature, talking with managers and scientists, and field visits, acreage estimates for several other stands were revised (Croatan, NC, Flomaton, AL, and Moody Tract, GA). The Mississippi stands were found to not be “old- growth”.

Fort McClellan (#2) was the only addition to the list. Other stands (see Additions?) may be added when sufficient data and field visits prove they are indeed old-growth forests. These include tracts in Alabama (Kaul Tract), Georgia (Berry College), Louisiana (Kisatchie NF), and Texas (Angelina NF).

The lack of geographical and ecosystem representation is alarming. Louisiana, Mississippi, South Carolina, Texas, and Virginia may have no old-growth remaining (but see Additions? above). Also, the lack of any Flatwoods, Piedmont, and Cumberland Plateau stands is troublesome. These points highlight the fact that our vision and understanding for all longleaf pine forest dynamics and benchmark conditions are based on a few clayhill and sandhill stands that are unique ecosystems of their own. This fact should encourage us to search harder to find these missing links.

Major threats to these remnants include: wildfire (58% of extant stands), non-native species invasion (25%), urban interface issues (25%), and uncertain management outlook (25%). Wildfire (including hastily planned prescribed fires) threatens the largest number, and its damage (large-scale mortality after fire) may be the greatest threat to the continued existence of these stands. Fire-suppression’s legacy of heavy fuel accumulations will be one of the most pressing issues in longleaf pine restoration management (see Kush et al. 1999b).

However preliminary, this list and information represents the best and most complete information known on old-growth longleaf pine stands. As is obvious, aside from a few stands, very little is known about these stands. This represents another void in longleaf pine research and conservation. Our next step is to add to the old-growth core collection at Auburn University’s Longleaf Pine Stand Dynamics Lab. This and more detailed surveys and projects should give science a better understanding on past Southeastern climates, disturbance factors, harvesting, and effects of ecological restoration on the pinelands landscape. The fact that so few old-growth stands remain should encourage quicker action to conserve these remnants - in the form of careful restoration, solidifying land tenure, and land purchases and/or easements. Let’s get to work!

ALABAMA (159 acres) 1. Flomaton Natural Area (Escambia Co.) International Paper Company; 58 ac; Clayhill site; Fire-excluded virgin stand, with trees exceeding 300 years. Aggressive ecological restoration program including prescribed fire and hardwood harvesting began

205

in 1994. Active education, restoration, and science programs. Threats: non-native species invasions, wildfire danger, urban interface. Many research projects completed and underway (see Kush et al. 1999b, Meldahl et al. 1999, Varner et al. 1999, and Kush et al. this Conference).

2. Fort McClellan (Calhoun Co.) Department of Defense (Army); 101 ac; Mountain Site; A collection of 12 stands, ranging in size from less than 1 ac to > 40 ac. In varied burning regimes, from annual to extremely fire-excluded. Contains the only known frequently burned old-growth longleaf pine stands in the Mountain Province. Located on a recently closed Army base, these stands face an uncertain near- and long-range future. An Environmental Assessment is underway to transfer the Fort to the US Fish and Wildlife Service to become the Mountain Longleaf National Wildlife Refuge. Threats: wildfire danger, urban interface, uncertain management future. Several past and current research projects (see Varner et al. 1999 and others). The most well-studied mountain longleaf pine forest.

FLORIDA(2132 acres) 3. Big Pine Tract (Hernando Co.) FL Fish & Wildlife Cons. Comm.; 420 ac; Sandhill Site; A large, infrequently burned tract in central Florida. Ages are on the young side for old-growth (max. ~ 160 years), and understory is in poor shape. Acreage estimate is probably too high. Active restoration program in place (hardwood harvesting and burning). Many longleaf pine forests – associated species present, including gopher tortoises, and many herbaceous species. One of the few stands in existence that is part of a large wildland. Threats: wildfire danger, non-native species invasions. Very little research, with only 1 survey project performed (Beckwith 1967).

4. Eglin AFB Stands (Okaloosa Co.) Dept. of Defense (Air Force); 1712 – 9000 ac; Sandhill Sites; The largest remaining old-growth stands, in various stages of burning (biennial to extremely fire-excluded). Acreage estimates abound, highlighting the need for a more systematic survey of the Base. Stands contain very old trees (> 400 years). Several stands have recently experienced mortality resulting from re-introduction of fire. Most pinelands flora and fauna are present. These stands face the least risk from urban areas, and represent the most ecologically viable old-growth longleaf pine landscape. Threats: wildfire danger. Research limited to reports and 1 recent publication (Provencher et al. 2000).

GEORGIA (2024 acres) 5. Greenwood Plantation (Thomas Co.) Private; 500 ac; Clayhills Site; Often called “The Big Woods”, this stand is one of the most awe-inspiring tracts in existence. For the past 50 years, managed under the Stoddard-Neel Selection Method (USDA 1995) of timber-wildlife management. The stand is frequently burned, contains many very large trees, and a diversity of plant communities. All native flora and fauna typical of pinelands can be found. Threats: uncertain management future No formal research has taken place in this stand. Several fire management-related notes have been published by Neel, Komarek and others related to Tall Timbers Research Station.

6. Moody Tract (Appling Co.) The Nature Conservancy; 320 ac; Sandhill-Clayhill Site; A recent (October 2000) purchase by The Nature Conservancy of Georgia. A very large, though somewhat fire-excluded stand, with trees exceeding 250 years. Most floral and faunal elements remain, including abundant wiregrass, gopher tortoises, and red- cockaded woodpeckers. In need of an active ecological restoration program. The large size of the tract (largest in Atlantic Coastal Plain), its inclusion in a large protected landscape, and its location in the lower Atlantic Coastal Plain make it unique in this list. Threats: wildfire danger No research projects or publications.

206 7. Wade Tract (Thomas Co.) Private (Tall Timbers Easement); 206 ac; Clayhill Site; The classic old-growth longleaf pine savanna. The stand contains an intact fauna and flora and majestic views of the pinelands landscape. Threats: urban interface issues Several past and current research projects on stand dynamics, fire ecology, avian ecology, and many others (see Platt et al. 1988, Engstrom and Sanders 1997, and many others originating from Tall Timbers Research Station).

8. Thomasville Plantations (Thomas Co.) Private (Tall Timbers & TNC Easements); ~1000 ac; Clayhill Sites A collection of privately-owned stands, in various states of conservation (many are under Easements with Tall Timbers Research Station and/or The Nature Conservancy). Most are frequently burned and managed for bobwhite quail and other pinelands fauna and flora. These stands, in combination with #6 and #7 comprise the largest and best collection of old-growth longleaf pine forests and savannas. Most are/were managed under the Stoddard-Neel Selection Method of timber-wildlife management (USDA 1995) that has maintained these tracts in a pristine condition. Conservation efforts have been and should continue to be placed here to preserve this magnificent, biologically significant landscape. Threats: uncertain management future, urban interface issues Several past and current research projects under the direction of Tall Timbers.

NORTH CAROLINA (320 acres) 9. Bonnie Doone Tract (Cumberland Co.) City of Fayetteville, NC; 160 ac; Clayhills site; A rare tract of “round timber” in the pinelands of North Carolina. Stand is fire-excluded and in need of a careful restoration program. Threats: wildfire danger, urban interface issues No past or present research projects or publications.

10. Boyd Tract (Moore Co.) NC Division of Parks & Recreation; 60 ac; Clayhills and Sandhills Site; An old (trees > 350 years) stand that had undergone fire-exclusion for ~ 80 years. Restoration efforts (burning and raking) are underway. Old turpentining scars are present on most trees, providing a unique aspect and cultural history. Threats: wildfire danger, urban interface, non-native species invasions. Several past and current research projects (see Gilliam et al. 1993 and others).

11. Camp Lejeune Tract (Onslow Co.) Dept. of Defense (Marine Corps); ~50 ac (20 ha); Sandhill Site; A poorly known sand ridge stand, frequently burned. No data and not visited. Threats: ? No specific research projects or publications found.

12. Croatan Ridge (Craven Co.) USDA Forest Service; 50 ac (20 ha); Sandhill Sites; A pristine sand-ridge on the southern border of the Croatan National Forest. The ridge is frequently burned and contains an extremely diverse herbaceous flora. The overstory is not very old (max. ages = 180 years), but more sampling is needed. Another old stand (Catfish Lake Bay Rims) is nearby, and may be included with more data and visits (pers. comm., Mike Schafale, NC Natural Heritage Program). The acreage is a conservative estimate, and a more exhaustive field survey is needed. Threats: none Surprisingly, very little research performed, with no available publications.

ADDITIONS? Kaul Tract (AL) Large mountain longleaf pine forest in Coosa County – lacking data and visits. Berry College (GA) Mountain longleaf pine forest near Rome with old trees – more data needed on acreage and conditions (see Birkhead and McGuire 1999). Calcasieu Tract (Kisatchie NF, LA) Potential 300 – 600 acres, making it the only old-growth longleaf pine stand in Louisiana, and potentially the largest west of the Mississippi River. Visits and data needed.

207 Cape Fear Stand (NC) – A privately held stand that needs visits and data. Boykin Spring Longleaf (Angelina NF, TX) Potential 90 acre stand – would be the only old-growth longleaf pine stand in Texas. Visits and data needed.

ACKNOWLEDGEMENTS Many thanks to the managers and owners of these stands for their far-sighted conservation and management of these resources and their assistance with this project, specifically – Foster Dickard & International Paper (Flomaton Natural Area, AL), many with US Department of Defense (Fort McClellan, AL; Eglin AFB, FL; Camp Lejeune, NC), FL Fish and Wildlife Conservation Commission (Chinsegut Nature Center, FL), Leon Neel (Thomasville, GA Plantations), The Nature Conservancy (Thomasville, GA Plantations and Moody Tract, GA), Tall Timbers Research Station (Thomasville, GA plantations), USDA Forest Service (Pringle Road, NC; Calcasieu Stand, LA), and NC Division of Parks and Recreation (Boyd Tract, NC). This compilation would not have been possible without the encouragement and knowledge of Leon Neel, Bruce Means, Julie Moore, Bill Boyer, Janisse Ray, Allison McGee, Mike Schafale, Carolyn Kindell and many other generous souls.

LITERATURE CITED Birkhead, R. and J. P. McGuire. 1999. Lavender Mountain old-growth longleaf pine forest. Pp. 59 –61 in: Kush, J. S. (comp.) Proc. of the 2nd Longleaf Alliance Conference. Engstrom, R. T. and F. J. Sanders. 1997. Red-cockaded woodpecker foraging ecology in an old-growth longleaf pine forest. Wilson Bulletin 109(2): 203-217. Gilliam, F. S., B. M. Yurish, and L. M. Goodwin. 1993. Community composition of an old-growth longleaf pine forest: relationship to soil texture. Bull. of the Torrey Botanical Club 120(3):287-294. Hunter, M. L. and A. S. White. 1997. Ecological thresholds and the definition of old-growth forest stands. Natural Areas Journal 17(4):292-296. Kush, J. S., J. M. Varner, and R. S. Meldahl. 1999a. Old-growth longleaf pine forests in Alabama: What have we done, what can we do? 5th Eastern Old-growth Forest Conf. Duluth, MN. Poster. Kush, J. S., J. M. Varner, and R. S. Meldahl. 1999b. Slow down, don’t burn too fast… got to make that old- growth last. Pp.109-111 in: Kush, J. S. (comp.) Proc. of the 2nd Longleaf Alliance Conference. Landers, J. L. and W. D. Boyer. 1999. An old-growth definition for upland longleaf and south Florida slash pine forests, woodlands, and savannas. USDA Forest Service GTR. SRS-29. So. Research Station, Asheville, NC. 15 p. McWhite, R. W., J. Furman, C. J. Petrick, and S. M. Seiber. 1999. Integrated natural resources transitional plan, Eglin Air Force Base, 1998 – 2001. Eglin AFB Jackson Guard, Niceville, FL. 229 p. Means, D. B. 1996. The longleaf ecosystem, going, going… Pp. 210-219 in: Davis, M. B. (ed.) Eastern old- growth forests: Prospects for rediscovery and recovery. Island Press. Washington, DC. Meldahl, R. S., N. Pederson, J. S. Kush, and J. M. Varner. 1999. Dendrochronological investigations of climate and competitive effects on longleaf pine growth. Pp. 265-285 in: R. Wimmer and R. E. Vetter (eds.). Tree Ring Analysis: Biological, Methodological, and Environmental Aspects. CAB International, Wallingford, UK. Platt, W. J., G. W. Evans, and S. L. Rathbun. 1988. The population dynamics of a long-lived conifer (Pinus palustris). American Naturalist 131(4):491-525. Provencher, L. and Seven others. 2000. Longleaf pine and oak responses to hardwood reduction techniques in fire-suppressed sandhills in northwest Florida. Forest Ecol. & Mgmt. 1-15 USDA. 1995. Record of decision – Final EIS for management of the red-cockaded woodpecker and its habitat on National Forests in the Southern Region. Vol. 1. USDA Forest Service R8-MB 73. 162 p. Varner, J. M., J. S. Kush, and R. S. Meldahl. 1999. Ecological restoration of an old-growth longleaf pine stand utilizing prescribed fire. Tall Timbers Fire Ecology Conference Proceedings. 21:216-219. Varner, J. M., J. S. Kush, and R. S. Meldahl. 2000. The Mountain Longleaf Pine Resources of Fort McClellan, Alabama: Final Report on their Status, Ecology, and Management Needs. Fort McClellan Directorate of Environment. 83 p.

208 UNDERSTORY SPECIES COMPOSITION OF OLD-GROWTH MOUNTAIN LONGLEAF PINE (PINUS PALUSTRIS MILL.) STANDS AT FORT MCCLELLAN, ALABAMA

J. Morgan Varner, III (Interdisciplinary Ecology Program, Box 118526, University of Florida, Gainesville, FL 32611-8526) John S. Kush (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849) Ralph S. Meldahl (Auburn University School of Forestry & Wildlife Sciences, 108 M. White Smith Hall, Auburn University, AL 36849)

ABSTRACT: Historically, frequent lightning-ignited fire perpetuated open canopied longleaf pine forests carpeted by the most diverse understory plant community outside of the Tropics. Only half of extant longleaf pine stands show signs of recent fire, leading to dramatic alterations in plant species composition and diversity. Pristine plant species composition data is available for several of longleaf pine’s distinct forested regions, a notable exception being its mountainous communities in Alabama and Georgia, collectively termed mountain longleaf pine forest. Understory vegetation was sampled in three old-growth mountain longleaf pine stands (two frequently burned, one fire-excluded) at Fort McClellan, a US Army post in the Blue Ridge physiographic province of northeastern Alabama. Over the spring, summer, and fall 1999 study period, 80 native species were encountered in sampling plots, representing 58 genera and 33 families. Understory communities in frequently burned stands contained a highly diverse herbaceous component, dominated by grasses, (principally Andropogon ternarius), asters (Coreopsis major, graminifolia, Helianthus microcephallus, Solidago odora, and others), and many legumes. Cover and species richness were both very high, comparable to other frequently burned longleaf pine forests. In the fire-excluded stand, only 23 species were encountered in sampling quadrats. Importance in this stand was dominated by woody species (89.3 % of total importance) including Vaccinium pallidum, Acer rubrum, Sassafras albidum, Carya pallida, and Smilax rotundifolia; all species typical of fire-suppressed mountain longleaf pine forests in the region. Non-native species were not encountered in sampling plots, another justification for the pristine nature of Fort McClellan’s mountain longleaf pine forests. Finally, vegetational, historical, and landscape evidence is presented to argue that the historic fire regimes in mountain longleaf pine forests were very frequent, perhaps with a 1 to 5 year return interval. This study provides the first quantitative information on pristine mountain longleaf pine forest composition, and should aid in identification, conservation, and restoration of this endangered ecosystem.

INTRODUCTION While most regions within the range of longleaf pine forests and savannas have been well-studied, the mountains have remained a relative mystery. The mountain longleaf pine forest ecosystem stretches over the Piedmont, Ridge and Valley, Cumberland Plateau, and Blue Ridge physiographic provinces of Alabama and Georgia. These forests are unique in that they: grow at high elevations (up to 1900 ft.), commonly experience ice and snowstorms, are located at the maximal distance from tropical storm origins, dwell in a topographically diverse area, and lie in a climatically distinct region (Craul 1965). In light of these intricacies, very little information has been available on basic species composition or the role of fire in this undulating landscape.

As with the remainder of the longleaf pine forest, severe reductions in acreage have occurred since European settlement. Less than 100,000 acres of mountain longleaf pine forest remain. Of this total, a little over 100 acres has been identified as old-growth. The remainder of the acreage consists of mostly fire-excluded and cut-over woodlands (notable exceptions in Talladega NF and Fort McClellan, AL).

The role of fire in this landscape has been variously interpreted, and the most recent paradigm is one of relatively infrequent burning (7-25 year return interval; Frost 1998). Skeen and others (1993) go much farther, attributing fire in this region to a Native American creation of recent origin. Today, indeed, fire is noticeably absent or much reduced over most of the mountain landscape. Fragmentation by roads and land- use changes are obvious culprits, but early lumbering and lack of large-scale grazing probably played a major role in rapid woody succession in the region. Lumbering removed the fuel-producing pine canopy. Reduced grazing in the region probably led to reduced settler burning for forage, a practice that maintained pinelands in the Coastal Plain for nearly a century after lumbering.

209 The composition of this forest type has also remained unclear. Historical descriptions were clear: a vast open-canopied savanna with scattered scrub oaks and carpeted with grasses and asters (Sargent 1884, Reed 1905, Harper 1943). Which species, in what quantity and relative dominance are questions that are important for identification, conservation, and restoration of this ecosystem.

To address the lack of study, this work was initiated in 1998. The primary objective in this study was to survey and quantify understory plant species composition of variably burned old-growth longleaf pine communities. These benchmark data and their information could then be utilized for identification, conservation, and restoration of this rare ecosystem. An additional objective was to establish permanent plots for long-term studies on these rare plant communities, data that are sorely lacking for this disappearing community.

METHODS Fort McClellan is a 19,000 acre U.S. Department of Defense Army garrison in Calhoun County, Alabama. Fort McClellan contains a large portion of Choccolocco Mountain and its spur ridges, all capped with Weisner quartzite. Climate is warm and humid, with mean annual temperature 62.50 F and annual precipitation averaging 52.2 inches. Soils were mapped as Rough Stony Land, Sandstone – a miscellaneous Typic Udult that is representative of most mountain longleaf pine forests (Craul 1965). Slopes are steep, commonly exceeding 50%.

Fort McClellan is unique in that it contains the finest remnant mountain longleaf pine forests, including 12 old-growth stands (Varner et al. 2000). At Fort McClellan, three old-growth stands with varying burning regimes (1-2, 2-3, and >25 years) were selected.

Within each intensive study area, plot centers and vegetation sampling quadrats (Bee Sting Hill 3 plots; Red- tail Ridge 4; and Caffey Hill 5) were installed. From each plot center, four 1.1 m X 1.1 m quadrats were placed in cardinal directions, each 5 m from plot center, making 12 quadrats at Bee Sting Hill, 20 quadrats at Caffey Hill, and 16 at Red-tail Ridge. All quadrats were sampled once during spring (April), summer (June), and fall (September) 1999. Within each quadrat, all species were identified and cover was visually estimated (1 to 100 percent). Since sampling took place over three seasons, all three (spring, summer, and fall) cover values were averaged for calculations. Relative frequency (RF) and relative cover (RC) were calculated for both sites. Importance values (IV) were calculated for individual species as the sum of relative frequency (RF) and relative cover (RC), for a potential importance value equal to 200.

RESULTS Seventy-seven understory species were encountered over the three sampling periods. Most species were represented at both frequently-burned stands (42%). Red-tail Ridge was the most species-rich (62 species) and had highest average cover values (85% cover). The two frequently burned stands had more forbs (47 – 34 – 8; RTR – CH – BSH) and grasses (5 – 5 – 2) than the infrequently burned stand. Numbers of legumes (Poaceae), asters (), shrubs, and tree species were similar between the frequently burned stands, but differed greatly from Bee Sting Hill. Individual sample species richness maxima were 22 – 19 – 10.

The burned stands were dominated by a typical pinelands flora. Herbaceous layers were dominated by Andropogon ternarius, Chrysopsis graminifolia, Coreopsis major, Euphorbia corollata, Helianthus microcephalus, Panicum commutatum, Pteridium aquilinum, Rhus copallina, and Solidago odora. Asters, grasses, and legumes were the most well represented families, together comprising 56 – 50 (RTR – CH ) percent importance in the two frequently burned stands, respectively. Herbaceous species, in general, were most important at both sites, accounting for 85 – 71.5 percent of total importance at the two sites. Seedlings of tree species were rare, together comprising only 5 – 4 percent of total importance at Caffey Hill and Red- tail Ridge. Shrubs were also rare, accounting for 23.5 and 11 percent total importance at Caffey Hill and Red-tail Ridge. No non-native species were encountered in any of the burned plot quadrats in any season, or at either of the two sites outside of sampling quadrats.

Bee Sting Hill, a stand typical of most mountain longleaf pine stands in the region (personal observations), was typified by a paltry herbaceous layer and heavy woody dominance. Several hardwoods (Acer rubrum, Sassafras albidum, Carya pallida, and other Quercus spp.) dominated the under- and midstory of the

210 infrequently burned stand. Vaccinium pallidum, a thicket-forming shrub, dominated, much as it does on many fire-suppressed mountain longleaf pine stands (Golden 1979, Maceina et al. 2000). Few grasses (2) or forbs (8) were found within the plots, or in the stand (Tables 1 and 2).

Table 1. List of most important species in three old-growth mountain longleaf pine stands at Fort McClellan, AL. Importance values were calculated where IV = Relative Frequency + Relative Cover. Both frequency and cover values were averaged over the three samples (SP, SU, and FA) in 1999. Form refers to plant form, either tree (T), shrub (S), vine (V), forb (F), or grass (G).

Site Fire History Top 10 Species (IV200) IV200 Form

Bee Sting Hill >25 Vaccinium pallidum Lowbush Blueberry 40 S Acer rubrum Red Maple 31 T Sassafras albidum Sassafras 29 S Carya pallida Sand hickory 18 T Smilax rotundifolia Bullbrier 16 V Diospyros virginiana Persimmon 12 T Quercus montana Rock Chestnut Oak 10 T Oxydendron arboreum Sourwood 7 T Quercus velutina Black Oak 6 T Nyssa sylvatica Blackgum 5 T Pteridium aquilinum Bracken Fern 5 F

Caffey Hill 2-3 Sassafras albidum Sassafras 20 S Andropogon ternarius Paintbrush Bluestem 16 G Coreopsis major Greater Coreopsis 13 F Rhus copallina Winged Sumac 12 S Solidago odora Sweet Goldenrod 10 F Pteridium aquilinum Bracken Fern 10 F Chrysopsis graminifolia Grass-leaved Aster 9 F Panicum commutatum Variable Panicgrass 9 G Euphorbia corollata Flowering Spurge 9 F Diospyros virginiana Persimmon 8 T Carya pallida Sand Hickory 8 T

Red-tail Ridge 1-2 Andropogon ternarius Paintbrush Bluestem 20 G Helianthus microcephalus Small Sunflower 13 F Rhus copallina Winged Sumac 9 S Pteridium aquilinum Bracken Fern 9 F Euphorbia corollata Flowering Spurge 9 F Stipa avenacea Blackseed Needlegrass 8 G Aster patens Late Purple Aster 8 F Aster undulatus Wavyleaf Aster 6 F Chrysopsis graminifolia Grass-leaved Aster 6 F Clitoria mariana Atlantic Pigeonwings 6 F Tephrosia virginiana Goat’s Rue 6 F Vaccinium arboreum Sparkleberry 6 S

211 Table 2. Summary table of understory plant species composition of three old-growth mountain longleaf pine stands with varying burning histories.

Bee Sting Hill Caffey Hill Red-tail Ridge Burning Regime >25y 2-3 1-2 No. species 23 48 62 Total number of woody species 12 9 10 Total number of forb species 8 34 47 Asters 2 13 12 Legumes 1 7 10 Total number of grass species 3 5 5 Quadrats with Bluestems (%) 8 75 100 Total number of exotics 0 0 0 Max. Species Richness (Spp. Quadrat-1) 10 19 22 Mean Species Richness (Spp. Quadrat-1) 4.5 12.3 13.5 Max. Cover Values (%) 98 98 105 Mean cover values (%) 45 81 85

DISCUSSION Fire-maintained pinelands are renowned for being extremely species rich (Peet and Allard 1993 and refs therein), and the two frequently burned stands at Fort McClellan were no exception. First, even with such limited sampling, 72 native understory plant species were encountered in the frequently burned stands. Single sample maxima of 22 and 19 species per 1.21 m2 quadrat are comparable to many sites listed in Peet and Allard (1993). These findings are significant, in light of the fact that data from mountain longleaf pine communities were not included in Peet and Allard (1993) or any similar work. Woody species dominated the fire-excluded stand, with forbs comprising only 44% of total importance. Finally, In contrast to other mountain longleaf pine community investigations (Golden 1979, Whetstone et al. 1996, Maceina et al. 2000), we found no non-native species in any of our 48 sampling quadrats. As with Coastal Plain longleaf pine forests, deviations the open, nearly monospecific canopy of longleaf pine with a patchy scrub midstory and a diverse grassy-herbaceous understory signifies alterations in fire regimes.

Using the same types of data as Frost (1998; vegetation and historical evidence), we argue that fire in mountain longleaf pine forests was frequent, potentially as frequent as Coastal Plain fire. First, the presence and dominance of communities dominated by longleaf pine, itself one of the most fire dependent species in North America suggests that frequent fire regimes must have been in place for an extended period. Historic descriptions and these contemporary findings of understory plant species composition (domination by a dense cover of shade-intolerant herbaceous species with few shrubs) further suggest pervasive and frequent historical fire regimes. Historical evidence for frequent fire regimes in mountain longleaf pine forests is voluminous. Reed (1905) estimated that longleaf pine forests dominated 87 percent of his large Coosa County mountain longleaf pine landscape.

Adjacent plant communities in mountain landscapes provide more support for frequent presettlement fire regimes. Pitcher plant bogs, orchid bogs, cane brakes, and Xyris swales are all located within the mountain longleaf pine landscape matrix. These associated communities all require frequent (1 to 5 year return interval) fire regimes for their maintenance and persistence, even in the mountain region (Hughes 1966, Folkerts 1982, ANHP 1994). As in the Coastal Plain, these associated communities are recipients of fires that originated in, and were carried by, the upland longleaf pine forests (Means 1996). Indeed, fire was (and at Fort McClellan, still is) so frequent in the mountains that longleaf pine communities were able to dominate cool, moist north-facing slopes (Reed 1905, Andrews 1917, Harper 1943, Peet and Allard 1993, Varner et al. 2000). Surely, to have had such an impact on the landscape, longleaf pine and frequent fire must have pervaded the region long before Native Americans settled the southeastern USA. For mountain longleaf pine forests to maintain a viable place in the future landscape, it will be necessary to better understand fire as both an evolutionary and management force.

212 ACKNOWLEDGEMENTS Richard Sampson, Chadwick Avery, Shawn Harrison, and Dan Spaulding provided field and laboratory assistance for this study. Discussions with Debbie Folkerts, Mark MacKenzie, Bill Boyer, and Doria Gordon greatly enhanced the quality of this paper. Additionally, the assistance provided by Directorate of Environment and US Fish and Wildlife Service personnel at Fort McClellan, notably Ron Smith, Gordon Horsley, and Bill Garland, was invaluable. Funding was provided by a US Dept. of Defense Legacy Fund, Cooperative Agreement No. USFS-SRS-33-CA-97019 between the US Army Directorate of Environment, USDA Forest Service Southern Research Station, and the Auburn University School of Forestry and Wildlife Sciences. The senior author wishes to thank the Interdisciplinary Ecology Program and Botany Department at the University of Florida for funding while this poster and manuscript were prepared.

LITERATURE CITED Alabama Natural Heritage Program (ANHP). 1994. Natural heritage inventory of Fort McClellan, Main Post. Alabama Natural Heritage Program. Montgomery, AL. 191 p. Andrews, E. F. 1917. Agency of fire in propagation of longleaf pines. Botanical Gazette 64:497-508. Craul, P. J. 1965. Longleaf pine site zones. Unpublished final report on file with USDA Forest Service Southern Research Station. Auburn, AL. 58 p. Folkerts, G. W. 1982. The Gulf Coast pitcher plant bogs. American Scientist 70:260-267. Frost, C. C. 1998. Presettlement fire regimes of the United States: A first approximation. Tall Timbers Fire Ecology Proceedings 20: 70-81. Tall Timbers Research Station, Tallahassee, FL. 460 p. Golden, M. S. 1979. Forest vegetation of the lower Alabama Piedmont. Ecology 60(4):770-782. Harper, R. M. 1905. Some noteworthy stations for Pinus palustris. Torreya 5:55-60. Harper, R. M. 1943. Forests of Alabama. Monograph 10. Geological Survey of Alabama, Tuscaloosa, AL. 230 p. Hughes, R. H. 1966. Fire ecology of canebrakes. Tall Timbers Fire Ecology Proceedings 5:149-158. Tall Timbers Research Station, Tallahassee, FL. 208 p. Maceina, E. C., J. S. Kush, and R. S. Meldahl. 2000. Vegetational survey of a Montane Longleaf Pine Community at Fort McClellan, Alabama. Castanea. 65(2): 147-154. Means, D. B. 1996. The longleaf ecosystem, going, going… Pp. 210-219 in: Davis, M. B. (ed.) Eastern old- growth forests: Prospects for rediscovery and recovery. Island Press. Washington, DC. Peet, R. K. and D. J. Allard. 1993. Longleaf pine vegetation of the southern Atlantic and eastern Gulf Coast Regions: a preliminary classification. Tall Timbers Fire Ecology Proceedings 18:45-82. Tall Timbers Research Station, Tallahassee, FL. 418 p. Reed, F. W. 1905. A working plan for forest lands in central Alabama. USDA Forest Service Bull. 68. 71 p. Sargent, C. S. 1884. Report on the forest of North America (exclusive of Mexico). US Department of Interior Census Office. Tenth Census Report. Vol. 9 612 p. Skeen, J. N., P. D. Doerr, and D. H. Van Lear. 1993. Oak-Hickory-Pine Forest. Pp. 1-33 in: Martin, W. H., S. G. Boyce, A. C. Echternacht (eds.) Biodiversity of the southeastern United States: upland terrestrial communities. Wiley. New York. Varner, J. M., J. S. Kush, and R. S. Meldahl. 2000. The Mountain Longleaf Pine Resources of Fort McClellan, Alabama: Final Report on their Status, Ecology, and Management Needs. Fort McClellan Directorate of Environment. 83 p. Whetstone, R. D., J. M. Ballard, L. M. Hodge, and D. D. Spaulding. 1996. Vascular flora of Fort McClellan, Calhoun County, Alabama. Whetstone Consulting, Inc., Anniston, AL. 153 p.

213 FIRE IN LONGLEAF PINE – A PRIMER

Dale Wade (USDA Forest Service, Athens, GA 30602)

ABSTACT: The Longleaf Alliance will soon publish a pamphlet discussing the use of prescribed fire in longleaf pine management. The target audience is the private landowner who wants to learn more about the reasons for, and mechanics of using prescription fire to enhance his/her land holdings. This project is currently in the rough draft stage. The poster contains the draft Table of Contents and some of the figures that are planned for inclusion. Suggestions on topics you would like to see covered in the text or in an appendix are encouraged and should be sent either to the Longleaf Alliance or to Dale Wade at [email protected]

214 RESTORATION AND MANAGEMENT OF LONGLEAF PINE FORESTS AND SAVANNAS IN BIG THICKET NATIONAL PRESERVE, TEXAS

Roy Mason Zipp (Big Thicket National Preserve, 3785 Milam, Beaumont, TX 77701)

ABSTRACT: The forests of Big Thicket National Preserve (97,000 acres) are recovering slowly from the lingering impacts of extensive logging and 75+ years of aggressive fire suppression. In fire-dependent upland forests, continued management with prescribed fire will eventually restore and maintain most of these pyric communities. However, in certain areas more intensive management efforts are needed to restore the Preserve’s remnant longleaf pine forests and savannas. This poster will illustrate several restoration successes. It will also present several proposed restoration projects including (1) conversion of a 71 acre slash pine plantation to a longleaf pine forest, and (2) restoration of a mixed hardwood pine thicket to a historic wetland pine savanna. Both of these projects are currently in a planning phase, and it is hoped that the poster will generate a lively discussion of innovative restoration techniques and strategies.

215