National Academies Press: OpenBook

Environmental Issues in Pacific Northwest Forest Management (2000)

Chapter: 8 A Framework for Sustainable Forest Management

« Previous: 7 Forest Management and Rural Communities in the Pacific Northwest
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 171
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 172
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 173
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 174
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 175
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 176
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 177
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 178
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 179
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 180
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 181
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 182
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 183
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 184
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 185
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 186
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 187
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 188
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 189
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 190
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 191
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 192
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 193
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 194
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 195
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 196
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 197
Suggested Citation:"8 A Framework for Sustainable Forest Management." National Research Council. 2000. Environmental Issues in Pacific Northwest Forest Management. Washington, DC: The National Academies Press. doi: 10.17226/4983.
×
Page 198

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

8 A FRAMEWORK FOR SUSTAINABLE FOREST MANAGEMENT /ntroc/uction Each of the goals iclentifiec3 by the committee (sustain viable populations of indigenous species, maintain properly functioning ecological processes, meet human needs for forest commodities, and satisfy cultural and aesthetic values) depends on the functioning of key ecological or ecosystem processes. Maintaining these processes is being adopted as the fundamental goal for sustainable forest management. Not all goals can be maximized concurrently; therefore, balancing partly incompatible goals through forest-management practices is the major challenge facing forest managers. Although much of the debate over forest management has focused on old-growth stands in forested landscapes and on effects of clearcut harvesting on biodiversity, sediment flow, and fisheries, forest-management practices have been in constant transition. Partial-harvest practices have become more widely used, newer harvest and replanting practices have lowered effects on soils and streams, and landscape analyses are being cleveloped to address issues of representation and connectedness of key habitat types. As they have in the past, forest-management practices will continue to change in response to new information and changing societal values. ELEMENTS OF FOREST MANAGEMENT Forest management has come to represent far more than simply logging 771

772 Pacific Northwest Forests anct silviculture aimed at fiber extraction. It is the variety of actions applied to forested landscapes to achieve the goals clescribed above. In more specific terms, the committee viewed forest management to comprise four elements: Allocation: the allocation of land to particular uses or combinations of uses. Rationing: rationing or scheduling use, such as levels of timber harvest, permitted recreation use, ant! bag limits on game and fish. Harvest: determining how to harvest or take forest products, which are often related to plans for the next cycle of management. Investment: investment in the productive resources of the land, including protection against wilcifire, insects, disease, and other threats; holding trees as Umber and resource capital; management practices such as fertilization or irrigation intended to increase value of forest products; and roads and infrastructure. Land A//ocation At the regional level, basic patterns of forest land ownership in relation to other uses such as development and agriculture have already been largely made. Congress has used ant] continues to use, its plenary authority over the federal lands to designate national parks, wilderness areas, wild anct scenic rivers, national recreation areas, ancl others. These decisions, while having a broad regional impact, have usually been made individually with little relation to each other or to other land categories, and with little attention to a regional scheme. The ecological consequences of regional land allocations among ownerships and jurisdictions have become increasingly clear. At the extreme, the so-called "checkerboard lands" have allocated uses and defined boundaries with no regard for ecological processes such as hydrology, natural disturbance, or wildlife behavior. Allocation of uses within particular classes of land ownership (e.g., federal, state, private, etc.) is generally macLe with little cognizance of ecological processes occurring at larger scales or of surrounding land uses and objectives. Although Congress has been specific in setting uses of some desig- nated federal areas, it has left allocating uses on the bulk of the national

A Framework for Sustainable Forest Management 773 forest and BUM land up to the land-management agencies. The Forest Service and BEM use a planning process for allocating land uses under broad multiple-use guidelines that give the agencies wide discretion in allocating federal lands to specific uses or mixture of uses. This planning process identifies streamside and scenic influence zones, special wildlife habitat, and intensive recreation use areas as well as timber management and harvest areas. Consideration is given is required to some extent- to landscape- or watershed-level consider- ations but for the most part, those plans are for federal lands only. Although they may be based partly on recognition of prior decisions on intermingled Private lands—em. patterns of previous locking in a 0 1 0 ' 1 1 · · · r ~ ~ ~ ~ ~ ~ · ~ A watershed—decisions in federal land-use plans only control actions on federal land. At the private and state levels, the Endangered Species Act has affected land avocation by means of habitat conservation plans, regulatory actions, and state programs and agreements. A notable example is the Oregon Plan for Salmon and Watersheds, which has as a goal to restore populations and fisheries to productive and sustainable levels that will provide substantial environmental, cultural, and economic benefits. Only on the larger private ownerships, usually those of forest industry firms, can private forest owners make decisions that encompass entire landscapes or watersheds. But, even the largest private forest ownerships in the re=mon are usually fragmented and intermingled with those of other owners. Each owner can allocate its forest lands to uses and intensities of use that depend on the owner's objectives. Land-use controls, such as those of the Oregon Land Conservation and Development Commission, may limit an owner's ability to change from forest to nonforest uses, but they typically do not require the owner to devote the land to particular forest uses. The major land avocations affecting forest land in the region are those in the Northwest Forest Plan. Those allocations of federal forests override previous allocations made in the planning process, but not those made by Congress. The allocations create a set of reserved old- growth forest areas, riparian reserves, and adaptive management areas on federal forests in the spotted owl-region, each with a specific set of management and use restrictions and guidelines. These old-growth reserves cover some 7million acres, about30% of the total area of federal forests in the spotted-owl region and roughly the same total area of

774 Pacific Northwest Forests forest that was already in congressionally withdrawn areas. in addition, the Northwest Forest Plan put about 9% of the federal forests ~ riparian reserves (pending watershed analyses) and 6% in adaptive management areas (FEMAT 1993~. Whether parallel allocations wiD be extended to Eastside federal forests as a result of current studies is uncertain at this time. Rationing Uses A basic forest-management and conservation method is to ration or schedule use over long periods to avoid depleting the ability of the forest to maintain productivity. This is common practice not only among government agencies, but also for forest industry firms in managing for timber production. The main management laws guiding the Forest Service and BEM require sustained yield (high-level annual or periodic flows) of all forest resources over time from federal forests. The Forest Service has interpreted this as meaning "nondeclining-even- flow" of timber from national forests. The idea of harvest scheduling is also captured in the concept of "bag limits" for fish and game, as wed as in the "carrying capacity" of rangelands. These are thought of as maximum sustainable levels of harvest in view of the availability of food and habitat for the species in question. For fish and wildlife, bag limits (population management) for all lands are set and enforced by the states. The level and timing of uses should be determined by an understanding of the capacity of forested ecosystems to produce desired goods and services sustainably as well as by market demand. Such "supply-side" rationing has been the stated goal on public and many private lands. Forest-management agencies and most forest-industry firms take a relatively long-term view and plan for relatively even and sustained annual timber harvests. Private forest owners other then fores/-industry firms are more likely to vary timber harvests based on market conditions and other factors, and they are unlikely to be committed to specific Tong- term harvest schedules. Timber-harvest schedules for a national forest or industrial forest are based on the size of the forest, the existing timber inventory, productivity of the land, expected investments in growing timber, and expected revenues, ah over a period of as much as several

A Framework for Sustainable Forest Management 775 decades. In the case of industrial forests, market strategies and the expected availability of timber from other owners may also play a role in sewing harvest schedules. Rationing or harvest scheduling is usually done for individual ownerships, or in the case of federal lands, for fairly large units such as nationalforests. Rarely is such rationing done across ownerships or jurisdictions. To the extent that an individual ownership or national forest encompasses a landscape or watershed or larger area, the rationing can be said to be done on a landscape/watershed or even a regional basis. The rotation age or time between successive cuts is an important element in setting timber harvest schedules. This could be the age at which the maximum total fiber is available for harvest, but today it is often the age at which forest stands are judged to be "financially mature." This is usually taken to mean the age at which the cost of holding trees for another period of time exceeds the increase in value that is expected over that time period. Financial maturity is influenced by the rate of increase in the volume of fiber or merchantable wood, the value of the trees, potential earnings from alternative investments (opportunity costs), and expected changes in real prices for wood. The importance of these factors varies depending on owner's objectives, expected markets, site productivity, and alternative opportunities for investing the money currently represented in the standing trees. No single rotation age fits all goals or circumstances, but the financial rotation age is generally considerably lower than the age at which the maximum total fiber is present. For many private owners in this region, planned rotation ages are 40-60 years, while the goal for public forests is generally longer. Over the past 60 years, the age structure of Pacific Northwest forests has been changed by logging, but generally with an eye ultimately to balancing the annual timber cut with annual growth over the long term. In the process, most of the old-growth was cut, the overall average age of the forest was reduced, and timber growth rates have increased. Although timber harvest schedules were not set jointly for public and private forests (except in the special case of the Cooperative Sustained Yield Unit in Shelton, Washington), expected harvests from federal and other public forests surely had some influence on harvest schedules for private forests. Scheduling decisions on one category of land ownership clearly

776 Pacific Northwest Forests influence demand and, therefore, decisions regarding the rationing of uses on other lands. The harvest schedules under the Northwest Forest Plan recognized the age structure of timber available for harvest. The plan left available a significant volume of timber outside of the late- successional and riparian reserves for planned harvesting. It was specifically planned to provide for most of the annual timber harvest until younger stands were oIcE enough to be cut. In addition, a consicler- able volume of young Ember was placed in the late-successional reserves that was to be available for certain kinds of commercial thinning. And habitat conservation plans (HCPs) implementecl uncLer the aegis of the Endangered Species Act that limit or eliminate harvest activities from particular public or private lancts will likely influence scheduling decisions on lands not included in the HCPs. Harvesting I egging is the most visible and probably the single forest-management action having the greatest impact on forests and their ecological processes. From the 1940s through the 19SOs, forest cutting on managed federal lands in the Douglas-fir region was accomplished with dispersed cutting units of approximately 40 acres in size. The resulting landscape was composed of many small forest patches and young plantations with extensive areas In altered microclimates, maximum edge habitat, and minimum interior habitat (Swanson and Franklin 1992) . This dispersing pattern was a stark contrast to the practices on private lands where progressive clear-cutting of large areas was employed. Among harvest methods, clearcutting in which all overstory trees are removed in a single operation has attracted the greatest attention. Reforestation of clear cuts is obtained by artificial planting, natural seeding from adjacent stances or frown trees cut during logging. Seed tree harvests involve removal of most of the overstory trees in a single operation, leaving behind singly or in groups a small number of seed producing togging is the most visible and probably the sing/e forest- management action having the greatest impact on forests and their eco/ogica/ processes.

Framework for Sustainable Forest Management 777 trees. In shelterwood cuts, the mature stand is removed in a series of cuttings that extend over a relatively short period of the rotation. An even-aged cohort of juvenile trees is thus encouraged in the shelter of uncut trees over the harvest period. A] of these harvest methods are associated with even-aged forest management (Smith 1986~. Uneven-aged forest management is used to maintain stands of trees with a full range of age classes up to the rotation age. it is generally used when the most commercially important species are relatively shade-tolerant and can grow under the shade of older trees. Trees are harvested as single, scattered individuals or in small groups at relatively frequent intervals throughout the stand and continuous regeneration of new trees is encouraged. Although the decision to harvest particular stands is generally a part of harvest scheduling for entire public and forest-industry ownerships, the choice of logging methods and associated regeneration actions (e.g., site preparation, planting, and vegetation control) is typically made on the basis of conditions specific to each site or stand. Even-aged management is suited to regeneration of relatively shade-intolerant tree species such as pines and Douglas fir, whereas single tree harvest and uneven-aged management work best for species such as hemlock and true firs which can establish and grow in partial shade. Even-aged systems are most widely associated with single-species stands; uneven- aged systems are more frequently used for multispecies management (Smith 1962~. Even-aged management initiated with clearcutting the forest overstory is generally used on industry and, at least until recently, on most federal lands on the Westside. Nonindustrial private forest owners are more likely to use some form of uneven-aged management or multi- entry even-aged management where overstories are partially removed (partial cutting) on the Westside, although clearcutting is still generally preferred. Partial cubing is occasionally used on all ownerships in other parts of the region. Typically, logging on steeper slopes (grade more than 30%) is done using cable systems to suspend logs and move them from the stumps to the loading deck. In especially sensitive areas logs may be transported by helicopter or balloon, although these techniques are comparatively expensive. On less-steep terrain, logs are moved with rubber-tired or tracked ground skidding machinery.

778 Pacific Northwest Forests The usual practice following harvesting on federal and state lands, forestry industry lands, and some nonindustrial private lands is to do some form of site preparation prior to planting or natural regeneration. Depending on the character of the forest, type of harvest, tree species wanted in the subsequent forest, and condition of the site following logging, site preparation can include broadcast burning of logging slash and undergrowth, piling logging slash with or without subsequent burning, scarification of the soil or killing brush with herbicides. Inasmuch as state forest regulations in all of the Pacific Northwest states except Montana require that logged areas be substantially regenerated to commercially desirable trees, site preparation is an economically integral part of timber harvesting. Silvicultural practices in the Pacific Northwest have been the subject of a great deal of scrutiny, and many new techniques and innovations are becoming evident. Among these are variable retention harvesting (Kohm and Franklin 1997), biodiversity pathways modeling, and increased rotation times. As use of new techniques becomes more frequent, they are being evaluated for their ability to achieve environ- mental objectives in managed stands and landscapes. Silvicultural approaches have been evaluated from the standpoint of providing wildlife habitat, plant species diversity, and aesthetic values (McComb et al. 1994; Hansen et al. 1995; Boyle et al 1997~. Franklin et al. (1997) reviewed alternative approaches to timber harvest that may be more accommodating to concerns about the conservation of biodiversity or key ecosystem processes. Although longer rotations certainly provide benefits in this regard, those benefits can be enhanced considerably if longer rotations are coupled to a "variable retention harvest system." This system is based on the strategy of retaining structural elements of the harvest stand woody debris, snags, seedtrees and shelterwood for at least the next rotation. Franklin et al. (1997) argue that retention of structural complexity serves three major purposes, "lifeboating," "structural enrichment," and "enhancing connectivity." Lifeboating refers to retained structures that provide refugia for biological elements that might otherwise be lost from the area. By structural enrichment they refer to the fact that the retention of structural complexity accelerates the eventual reinvasion of harvested stands by species lost at harvest (e.g., Carey 1995~. Retention

A Framework for Sustainable Forest Management 779 of complexity diminishes the extent of fragmentation of habitat created by harvesting by enhancing among stand connectivity on managed landscapes. Implementation of specific prescriptions for harvest retention is clearly limited by existing knowledge of the importance of particular structural elements in the diversity of Pacific Northwest forest types. That said, FrankI~net al. (1997) ar~emuchinformation does exist upon which to inaugurate such programs and that they should in any case be coupled to carefully designed monitoring programs and an adaptive management philosophy. Investment Forest-management investments include investments in protection (from wildfire, insects, diseases, and other depredations), in various management practices (e.g., planting trees), in infrastructure such as access roads and recreation facilities, and in developing management information. The major investment is usually in the trees themselves, whether they are held for timber production or for other purposes, and it increases with age and size of the trees. As a result, most forest management is very capital intensive. Federal forests typically carry much heavier timber inventories—and thus have greater investments in timber—than private forests in the region. Timber inventories (on a per acre basis) on federal lands are often two to three times those for privately managed lands. This has been identified by some as an example of "inefficient" management of federal timber (CIawson 1976), but different management objectives may account for the differences. Forest-management investments are often thought of as applying only to growing and using timber, but investments are also required to provide nonwood products such as wildlife. Such investments include the opportunity cost of holding timber inventories to protect riparian and special wildlife habitat, to provide for landscape and other amenities that are attractive to both active and passive users of the forest, and to provide for the production of such miscellaneous products as Pacific yew and mushrooms. And just as there are opportunity costs in holding valuable timber off the market in order to get these

180 Pacific Northwest Forests nontimber benefits, there are opportunity costs associated with losing nontimber benefits In order to harvest timber and get its economic value. Those costs apply to both public and private forest owners. Examples of forest investment Postharvest Site Preparation. The objectives of site preparation include eliminating logging slash and other debris, reducing competition, modifying animal habitat, preparing a mineral seemed, mitigating compaction, improving drainage, creating a more favorable microsite for seedling establishment, and controlling disease. Depending on the character of the forest, type of harvest, tree species wanted in the subsequent forest, and condition of the site following logging, site preparation can include broadcast burning of logging slash and undergrowth, piling logging stash with or without subsequent burning, scarification of the soil, killing brush with herbicides, and application of fertilizers. Reproduction. The next stand of trees may be naturally seeded from nearby trees or planted with seedlings, in which case reproduction costs are quite low. Naturally seeded sites typically have very high stem densities and often require thinning at various stages in stand development. As stands develop in stature more of the thinned material can be sent to market, compensating for thinning costs. in many areas, particularly on large cuts or where natural seed rain is limiting, stands are regenerated by artificial planting. In many areas genetically improved seedling stock derived from parent trees selected to favor rapid growth or desired wood qualities is favored (Farnum et al. 1983~. Planting of trees at Tower densities significantly accelerates the establishment of the stand in the first five years after cutting. Intermediate SiZviculturaZ Investments. During the growth of a forest stand, a variety of treatments may be applied aimed at improving the existing forest stand, regulating its growth, or providing early financial Such treatments include pre-commercial and commercial thinning, improvement cuttings, pruning, release operations, and fertilization. Plantations and naturally-seeded stands typically have stem densities significantly higher than wid survive to maturity or that are desirable for optimal tree growth. A proportion of the trees in the overstory are cut at an early age (typically S-12 years) to thin tree density returns.

A Framework for Sustainable Forest Management 787 and accelerate growth on the remaining trees. Historically, Dinned trees have been left on site to decompose, but as stands mature some of the thinned material may be merchantable. Pruning of the lower tree branches at a young age is used in some cases to enhance future log value by providing knot-free wood. As in agricultural crops, research has shown that augmenting the soil with fertilizer can significantly increase tree growth (Wild and Breeze 1981~. Most commonly, nitrogen fertilizer in the form of urea pellets is applied from the air to the site approximately every 5 years. No other type of fertilizer is widely used. Where insect pests present serious risks, pesticides may be applied from the air. At nearly ad stages of stand development protection from wildfire is practiced, particularly in more arid regions. This may take the form of investment in personnel and machinery necessary to suppress fires or maintain protective structures such as fuel breaks and fire lanes. Increasingly, emphasis is being placed on the management of fuels by means of prescribed fire or brush removal. Grave} and dirt roads are the primary delivery system for logs from forest to mill. The vast forest road network that currently exists in the region represents a major technical and financial commitment of federal and state agencies and private landowners. Positive and negative incentives to forest investment Investment decisions are made in response to incentives and disincen- tives. Incentives might include increases in timber prices or the availability of tax breaks or cost-sharing by the public. Disincentives, such as uncertainty about the future, can reduce investments. For example, early in this century, perceived and real risks associated with wildfire and forest pathogens and pests were strong disincentives for forest investment on private lands (CIawson 1979~. Technological advances such as fire-suppression techniques and pesticides provided opportunities to reduce such risks and incentives for forestry invest- ments across a range of ownerships. Ironically, some of these technolo- gies (e.g., fire suppression) have led to landscape changes that have increased risk and uncertainty and complicated investment options in areas such as the Eastside of the Cascades. Positive and negative

782 Pacific Northwest Forests incentives are influencing forest-management investments in the Pacific Northwest. The reduction in federal timber harvests as a result of the Northwest Forest Plan is leading to increases in timber prices, which in turn lead to more intensive management of private forests for wood production. Uncertainty about Endangered Species Act requirements is apparently creating a perverse incentive for private owners, leading to premature timber harvesting and reduced management. In particular, uncertainty over whether disruption of spotted owl habitat on private land is prohibited is reportedly leading some private owners to eliminate such habitat by loggingbefore presence of spotted owls canbe verified. Others, mainly large landowners, have had "habitat conservation plans" accepted by the U.S. Fish anct Wildlife Service. For example, In return for some relaxation of restrictions on harvesting timber near spotted owl nests, the landowners agree to Tong-term habitat improvements (Weyerhaeuser 1994~. That kind of a response involves substantial investments not only in preparing materials for the permitting process, but in the modifications of timber management programs. The latter costs include the opportunity costs of postponing timber harvests and reducing potential timber yields, and the direct costs of using more costly management and logging practices. Investments on private forests are guided largely by the marketplace—expected costs and returns—as bounded by regulations and supported in some cases by public incentives. To some extent and in some places, these may be guided in part by arrangements among stakeholders, such as the Timber-Fish-Wildlife agreement in the state of Washington. Such ~nveshnents can be viewed as conforming to regional Or landscape level goals. On public forests, especially federal forests, investments are much less responsive to market incentives. Nevertheless, costs, including the opportunity costs of reserving land for specific purposes, are a vital issue. MEET/NO THE GOALS OF SUSTAINABLE FOREST MANAGEMENT During the first decades of this century, principles of "sustained yield" (Pinchot 1907) and "ecological forestry" (Toumey and Korstian ~ 937) were central as forest managers in North America began to appreciate

A Framework for Sustainable Forest Management 783 the need for and value of reforestation and planned scheduling of management activities. As demands for forest resources increased and became more complex, "multiple-use forestry" became the buzz-phrase of choice, particularly on public lands. Today, we speak in terms of "sustainable forest management" or "ecosystem management," recognizing not only the complexity of forest-management goals, but the complexity of the systems necessary to meet them. With 100 years of experience, forest managers have become adept at managing at least one aspect of a biological system according to a well- defined set of scientific principles and restricted economic and ecological objectives. The forest-management system has reasonably mastered the skills of managing over large spatial scales and learned routinely to make decisions and investments that will be realized only in temporal time scales of fifty or more years. Though management objectives may not always have been targeted correctly, the forest- management system as a whole has adapted to new knowledge and needs as they have become understood (Swanson and Franklin 1992~. The evolution to management for sustained yield in the 1950s is now followed by transition to ecosystem management in the 1990's (Salwasser 1990~. Learning has not always been smooth, and changes in forest management have often been fostered by public controversy (Chapman 1931; Swanson and Franklin 1992~. Management agencies often lack systematic plans for learning, which should include prioritized listings of identified uncertainties, methods for reducing important and tractable uncertainties, procedures for evaluating existing actions, and mechanisms for retaining new knowledge in the memory of the institution (Hilborn 1992; Christensen et al. 1996~. Incorporating large- scale and long-term ecosystem considerations in forest-management decisions generally requires changes in the behavior of the institutions the market, land ownership, laws and regulations, agency responsibilities—that frame these decisions. Institutional barriers to learning can limit our capacity to change (Senge ~ 990; Lee 1993~. The existing institutional framework is unlikely to be responsive to some of these challenges. The following section presents a framework for sustainable forest management, with particular reference to the challenges discussed in earlier chapters. The means by which forest management can be made adaptable and accountable in the face of inevitable uncertainty and

I84 Pacific Northwest Forests surprise is discussed. Finally, the challenges and opportunities for resolution of conflict and implementation of management policies across spatial scales ranging from individual ownerships to the entire region are considered. A FRAMEWORK FOR SUSTAINABLE FOREST MANAGEMENT Over the past several years, sustainable forest management or ecosystem management has been proffered as the fix for the variety of natural resource challenges facing the Pacific Northwest and elsewhere (e.g., FEMAT 1993; SAF 1993; SIocombe 1993; Gordon 1994; Grumbine 1994; Christensen et al. 1996~. However, the need to manage in the context of entire ecosystems rather than focusing only on selected parts of ecosystems has been recognized in the United States for more than half a century (SIocombe 1993, Grumbine 1994~. There is still much confusion about what ecosystem management means: to some, it is leaving more snags and logs in clearcuts, to others it is mimicking natural patterns at the landscape scale, and to yet others it is a comprehensive approach that recognizes structures, processes and interrelationships that span a wide range of scales from global to microscopic. AD commentaries on this subject have emphasized the importance of reconciling the complex human needs and interests as components of virtually ah ecosystems, while recognizing that sustained provision of natural resource commodities and amenities depends on the maintenance of biological diversity and ecosystem processes. Key elements of sustained forest management or ecosystem management (Grumbine 1994; Christensen et al. 1996) are described below.) Operational Goals Goals should be formulated so as to provide benchmarks for measuring the success of management policies and practices. Operational goals Adaptive management and monitoring are usually included in lists of key elements of ecosystem or sustainable forest management but are here discussed under the heading "Making Management Adaptable."

A Framework for Sustainable Forest Management 785 pay special attention to ecosystem structures and processes needed to sustain provision of natural resource goods and services. The goals of forest ecosystem management cannot be stated simply in terms of the commodities or amenities we would like to extract from such systems. Rather, management must acknowledge the importance of basic ecosystem structures and processes that sustain natural-resource goods and services. Given the complexity of ecosystem structure and processes, the setting of operational goals is by no means a simple matter. Nevertheless, the elements of ecosystem functioning discussed in Chapter 4 provide a framework for setting such goals. Managing in Context and Across Scales of Space and Time It is in the context of the unique circumstances and objectives of particular places that ecological principles are successfully applied to practical problems. The spatial and temporal context for management decisions should match the scales of ecosystem processes critical to sustainability. The connectivity of the landscape and the reality that actions at one location may influence ecosystem properties and processes elsewhere should be an integral part of management planning. Strategies must be developed to reconcile the incongruity of political jurisdictions and ownership boundaries with ecosystem processes. Issues of scale head almost everyone's list of forest ecosystem management objectives, and for good reason; successfully achieving any set of management goals requires that each goal be addressed at appropriate spatial and temporal scales. For example, sustaining viable populations of many species requires maintaining sufficient habitat at the regional scale. It also requires a long-term view that foresees inevitable habitat loss and provides for buffering and replacement. Sustaining the potential for sites to produce wood requires · A landscape perspective because landscape patterns influence the spread of disturbances. · A local ecosystem perspective because the structure of ecosystems influences their stability. · A microscopic perspective because nutrient cycling, hence forest

786 Pacific Northwest Forests productivity, requires healthy populations of microorganisms and microinvertebrates. ~ A temporal perspective because actions that increase tree growth in the short run could in some cases decrease it in the long run. · A global perspective because some factors that have a direct bearing on tree growth, such as climatic fluctuations and thinning of the ozone layer, are global processes. · A perspective on complexity, or the realization that the dynamics of nature emerge from interactions between management and numerous natural processes occurring within and across these various scales. Although forest management has commonly dealt with relatively large scales in making decisions about land allocation, management of multiple use lands, and in particular silvicultural strategies, have focused primarily at the stand {ever. In effect, the forested landscape was divided into two distinct parts: reserves and areas managed primarily for wood fiber (Perry 1994~. Many ecologists and conservation biologists have concluded that sustaining biological diversity in the long-run is not possible solely in a system of reserves (Noss 1983; Walters 1986; Perry and Maghembe 1989; Maser 1990; Norse 1990; Franklin 1991a; Salwasser 1991; Kessler et al. 1992~. That recognition has at least two implications for the scale at which forest-management decisions should be made. First, reserves must be located using criteria in addition to their aesthetic worth and low commercial timber value. Second, stand management, and especially silvicultural decisions on multiple use lands must consider goals and context at larger scales. The fact that a growing number of subregional management plans (i.e., smaller scales than FEMAT and PACFISH) cover spatial scales ranging from tens of thousands to hundreds of thousands of acres is encouraging. APPROACHES TO MANAGING FOR D/VERS/TY: GENERAL CONS/DERAT/ONS A primary challenge for forestry in the Pacific Northwest-indeed for forestry throughout the worId-is translating the guiding principles embodied in ecosystem management into specific practices on the ground and developing tools to assess how successfully goals are being

A Framework for Sustainable Forest Management 787 met Considerable progress is being made along these lines and, although the science and art of managing for sustainability is in its infancy the Pacific Northwest has been and remains a leader in developing and proving new techniques. A primary cha//enge for forestry in the Pacific Northwest-indeed for forestry throughout the wor/~-is trans/ating the guiding princip/es emboc/ied in ecosystem management into specific practices on the ground and deve/oping tools to assess how successfu//y goals are being men As discussed earlier, one of the primary lessons of modern ecology is the seamiessness of interactions in space and time. One implication of this is that landscapes and regions must be considered as interactive wholes rather than simple collections of individual stands. Nevertheless, when formulating management strategies at the level of landscapes and regions, it remains useful to distinguish land-use types that play different functional roles and that may or may not be tied in perpetuity to a fixed area. In what they called the triad approach, Seymour and Hunter (1992) distinguished three land-use types that, in theory, could coexist at some level within a region without compromising the goal of sustaining biological diversity: (a) reserves, (b) lands managed using "new forestry" techniques (i.e. where maintaining diversity and ecosystem function takes precedence over maximizing timber production), and (c) intensively managed lands (where the emphasis is on producing fiber). From the standpoint of maintaining biological diversity, the triad approach (or any other that recognizes different land-use types) reflects the fact that not every piece of ground must function as suitable habitat all the time to maintain viable populations. The key scientific questions center on the proportion and spatial arrangement of the three types that give a reasonable probability of maintaining diversity through time, the time requirement implying that dynamics-disturbance and recovery-must be taken into account. This is the approach proposed in FEMAT {19931 and adopted in the Northwest Forest Plan. On the other , ~ _ c . ~ ~ ~ ~ , . hand come ~~olosJical functions must be sustained on every piece of ground, especially those related to soils, nutrient cycling, and the interactions between land and water that regulate hydrologic flows and produce clean water.

788 Pacific Northwest Forests The biological crisis in the Pacific Northwest can be traced directly to a severe imbalance in land-use types. During the latter half of the 20th century, intensive forest management spread to cover a majority of the regional landscape. With the possible exception of the northern Rockies (~daho and western Montana), which have a high proportion of area in designated wilderness, the amount of area in reserves was grossly inadequate to sustain old-growth dependent species. Through implementation of plans developed by FEMAT, this imbalance is being redressed in the range of the northern spotted owl. Change is also occurring on federal lands outside the range of the northern spotted owl, especially in eastern Oregon and Washington, albeit more slowly. Throughout the region, but especially outside the range of the northern spotted owl, the degree to which changes are adequate to sustain ecological processes and old-growth dependent species remains to be seen. The interagency scientific analysis of the interior Columbia Basin recommended a management option for federal lands that emphasizes "active" management across the landscape rather than an increased commitment to reserves, arguing that was the best way to ensure restoration and maintenance of ecosystemhealth (USES 1996~. Managers appear to be left with virtually complete discretion about what form active management might take. The basic challenges for federal lands throughout the region are to (a) ensure a reasonable probability that the combined area in reserves and forests managed for ecological values is sufficient to maintain ecological functioning and biological diversity, and (b) address the scientific questions necessary to accomplish that successfully. With regard to reserves, the key questions center on how many, how large, where, and the degree of management needed to restore and maintain forest health. With regard to lands managed with an emphasis on ecology rather than production, the key question is what that means in terms of practices. These issues, especially the latter, are not restricted to public lands. With the growing influence of certification of forest products as having come from forests that have been managed in an environmentally suitable manner and general public awareness over environmental issues in forestry, private landowners are increasingly exploring management strategies that balance ecology and economics, a middle ground in which both objectives are served to some degree. In the remainder of this section we briefly discuss current efforts and ongoing issues related to translating the principles of ecosystem management to the ground.

A Framework for Sustainable Forest Management Reserves 789 Reserves are areas "managed primarily to maintain (or restore) the natural processes and conditions present prior to European settlement" (Aber et al 2000~. Reserves typically exclude activities such as timber harvest and road building, but may allow management aimed at restoring or maintaining desired conditions. In some cases, limited timber harvest may be permitted if deemed necessary to Tower risk of catastrophic disturbances. The science of reserve selection and placement has been recently reviewed by Noss and Cooperrider (1994) and Schwartz et al. (2000~. Recent models developed specifically for spotted owls include Andersen and Mahato (1995) and Hof and Raphael (1997~. It is beyond the scope of this report to deal with this topic in detail, however some specific issues are important to address, particularly the current debate about whether reserves are a wise strategy for maintaining diversity in the long run. The literature on the topic points out that the majority of scientists with expertise in this issue believe adequate and weB- distributed reserves are essential components of any conservation strategy (e.g. Noss ~ 983, Harris 1984, FEMAT 1993, Aber et al 2000~. A contrarian view has been expressed by some forest managers and academic foresters, who argue the best way to protect habitat in the long run is active management, including timber harvest. This view was expressed in greatest detail in the Report on Forest Health in the United States, written by a panel chartered by Congressman Charles Taylor (Oliver et al. 1997), and the approach was recommended by the Interior Columbia Basin Ecosystem Management Project (USFS 1996~. Propo- nents of active timber harvest on all or most of the landscape argue their approach reflects current ecological thinking, which recognizes nature as inherently dynamic, from which it follows that attempts to freeze any particular set of conditions are futile and misguided. Only through active timber harvest can stands be maintained in a healthy condition that continues to supply habitat (Oliver et al. 1997~. This view recently was critiqued by a pane] of the Ecological Society of America (ESA), which disagreed strongly with the conclusions of Oliver and colleagues (Aber et al. 2000~. This committee concurs with the ESA panel, which based its conclusions on three factors: · Although current ecological thinking indeed recognizes nature as

790 Pacific Northwest Forests dynamic rather than static, what this implies for conservation is not that reserves are a bad strategy, but that reserve design must account for dynamism in nature, largely by (a) ensuring enough area in reserves to subsume natural disturbance and recovery processes (e.g. FEMAT 1993), and (b) designing reserves so as to reduce risk of catastrophic distur- bance spreading from beyond their boundaries, which generally, though not necessarily always, means relatively large reserves well buffered at their boundaries (Noss 1983; Andersen and Mahato 1995; Schwartz et al. 2000~. · There is little evidence that managed stands are healthier than unmanaged stands. In fact, quite the contrary, experience both within and without the Pacific Northwest shows stands managed for timber are more susceptible to crown fires, pathogens, and insects than the natural forests they replaced (Huff et al. 1995; Schowalter et al. 1 997; Perry 1998; Aber et al. 2000~. · Even in the eventuality that forests could be logged so as to maintain habitat for the entire suite of old-growth dependent species, reserves still play a vital aesthetic and spiritual role for humans. A well-designed and properly managed reserve system, in which appropriate levels of management may play a legitimate role is an important feature of forest management. togging to improve Forest Health A number of forests in the Pacific Northwest that developed after clearcutting (moist forest types) or high-grade logging coupled with fire exclusion (dry forest types) are severely overstocked, a situation that was apparently uncommon before the arrival EuroAmericans (Mutch et al. 1993; Tappeiner et al. 1997~. Overstocking increases fire hazard in moist forest types west of the Cascades crest, and creates a variety of problems in dry forest types of the Klamath province and throughout the interior, including increased susceptibility to crown fires, insects, and pathogens (Perry 198Sb; Mutch et se. 1993~. Where overstocking threatens forest health, some careful logging may be appropriate even within reserves. FEMAT (1993), for example, proposed thinning in overstocked young stands within reserves to reduce fire hazard. Thinning understor,v trees within old-growth ponderosa pine forests has

Framework for Sustainable Forest Management 791 been proposed to reduce the risk of crown fire and insect infestation (e.g., Mutch et al. 1993), although many biologists and environmentalists are skeptical that benefits outweigh the environmental risks associated with logging. Togging to improve forest health is a complex issue that etudes simple generalizations. Thinning overly dense stands undoubtedly improves individual tree vigor, which In turn improves tree resistance to bark beetles, especially among the pines. Thinning may or may not reduce fire hazard, depending on which trees are dinned and how logging slash is treated. Thinning that lowers average stand diameter may well increase susceptibility to crown fires (Perry 1995a), and opening the stand too much could dry fuels and increase flammability. Experience from northern California shows partially logged stands in which logging slash is left untreated are more susceptible to crown fires than uniogged stands (Weatherspoon and Skinner 1995~. On the other hand, underthinning removes fire ladders that propagate flames into the crown of overstory trees, and hence reduces the chance of a ground fire becoming a crown fire. To successfully achieve the objective of reducing fire hazard, ah these factors must be taken into account and protocols developed for deciding which stands should be thinned and how much. Potential negative affects of logging should also be considered. Given the well-documented environmental problems associated with roads (Trombulak and Frissell 2000), logging to improve forest health would not seem justified if it required building roads into roadless areas. Issues of habitat must be considered. Some animals, notably goshawks and ungulates, require closed canopy forests in at least a part of their range and could be negatively impacted by widespread heavy thinning. In sum, judicious logging may be a good too! to improve the health of some forests, but each situation should be evaluated on its own merit, and operations planned carefully to ensure the cure is not worse than the disease. Management for Complexity and Diversity Much of human management is focused on ecosystem "simplification" or techniques to focus productivity into those elements of particular human interest. However, much research in ecology and forest management has shown that biological diversity and structural

792 Pacific Northwest Forests complexity of ecosystems are critical to such ecosystem processes as primary production and nutrient cycling. Complexity and diversity also impart resistance to and resilience from disturbance and provide the genetic resources necessary to adapt to long-term change. Three management strategies are key to the conservation of biological diversity. First, management practices for any one species must recognize that suitable habitat encompasses ah of the other species and system processes upon which that species depends. Second, area of habitat required to sustain viable species populations must be suffi- ciently large to buffer inevitable variations in population size through time. Third, as discussed above, a landscape and regional approach to the distribution of reserves and connections among them is critical. Variability and Change Forest ecosystems are constantly changing. Natural disturbances such as fire, windstorms, insect and pathogen epidemics, and floods are ubiquitous and, in many cases, critical to the maintenance of key ecosystem processes. Management determined to "freeze" ecosystems in a particular state is futile and unsustainable. Nowhere is the importance of var~ab~ty and change better demon- stratecl than with natural disturbances and the processes of ecological succession that derive from them. As discussed in Chapter 4, that issue is critical in the dry forest types of the Eastside (Mutch et al. 1993; USFS/BEM 1994; Hessburg et al. 1993; Sampson and Adams 1994~. Fire hazard is high, outbreaks of defoliating insects and, in some places, bark beetles have been more severe than at any time within the recorded past (e.g. Anderson et al. 1987; Wickman et al. 1992; USFS/BLM 1994), and the stage is set for severe pathogen problems (Hessburg et al. 1994~. There is consensus that the current poor state of those forests can be traced in part to decades of fire exclusion, high grade logging of ponderosa pine and western larch, and overgrazing (Anderson et se. 1987; Mutch et al. 1993; Covington et al. 1994; see Chapter 4~. Fire exclusion in many Eastside areas has resulted in large-scale conversion from a landscape dominated by old-growth, largely ponderosa pine, to forests that are younger, much denser, and domi- nated by tree species susceptible to western spruce budworm, ~ . . ... ~ .

A Framework for Sustainable Forest Management 793 Douglas-fir tussock moth, and several root rots. Reduced habitat (e.g., large dead wood and old-growth forests) for the natural enemies of defoliating insects may have exacerbated outbreaks (Torgersen et al. 1990; BuD and Holthausen 1993~. The combination of densely stocked young trees and (in many stands) large amounts of dead fuels resulting from insect outbreaks has significantly increased the probability that what in the past would have been a gentle ground fire will now become a severe crown fire. The following steps will be critical to dealing with this situation: thinning to reduce tree density; reintroduction of frequent ground fires; restricting any harvest in remnant old-growth stands to Winning understory trees; protection of any mature or old-growth ponderosa pine trees; maintenance of coarse woody debris as habitat for natural enemies; and longer rotations to restore the landscape to a high proportion of large, fire-resistant early successional tree species (primarily ponderosa pine) (Perry 1988a, Mutch et al.1993, Hessburg et al. 1993, Torgersen 1993, Arno et al 1993, Bull and Partridge 1986, Covington and Moore 1994, Henjum et al.1994, O'haughlin 1994, Oliver et al.1994, USFS/BT~M 1994~. However, the areas affected are quite large, and the remedial prescriptions are in some cases very expensive. Human development and building on fire-prone landscapes has increased the financial liability and risk to human life associated with catastrophic events that might occur as a consequence of extensive fuel accumulations. At the same time, such development severely constrains the variety of management interventions that may be used to remedy the situation. Prescribed fire in heavy fuel areas close to homes or other structures is itself risky and expensive and must be applied on a limited scale Furthermore, the use of prescribed fire is limited in some places owing to the effects of smoke on air quality near already polluted urban centers. Uncertainty and Surprise Three kinds of uncertainty are inherent in the management of forest ecosystems (Hilborn and Mange} 1997, Christensen et al. 1996~. First, uncertainty derives from unknowable responses that result from the complex and ever changing character of ecosystems. Examples of such

794 Pacific Northwest Forests uncertainty include ecosystem responses to climate change, rare events such as volcanic eruptions, or the cumulative effects of multiple environmental changes (e.g., the multiplicity of factors influencing populations of migratory fishes in forested landscapes). The fact that our forest ecosystems are being stressed in ways that may be unique with respect to their evolutionary history increases this kind of uncertainty. Uncertainties of this kind are difficult to eliminate or reduce, but Weir magnitude and relative importance can be estimated. Second, uncertainty arises because of limitations in our knowledge base or models of ecological and social systems. Uncertainties of this sort are being reduced by increased research, although extension of such information across scales of time and space represents a significant challenge (Levin 1992~. Poor data quality, sampling bias and analytical errors generate the Bird category of uncertainty. Managers and decision makers must work with scientists and data managers to determine an acceptable level of decision error. Public understanding and education on this matter is critical. Just as managers cannot claim perfect knowledge, the public should not expect it. Public understanding of the nature of uncertainty in our science and management, as well as the potential risks from surprises is a necessary prerequisite to informed involvement in decisions affecting the management of natural resources. Institutional barriers to learning can limit our capacity to reduce uncertainty (I ee 1993~. For example, management agencies often lack systematic plans for learning, which should include prioritized listings of identified uncertainties, methods for reducing important and tractable uncertainties, procedures for evaluating existing actions, and mechanisms for retaining new krlowI- edge in the memory of the institution (Hilborn and Mangel 1997, Christensen et al. 1996~. Dealing with these barriers is critical to the accountability and adaptability of management. Humans as Ecosystem Components Humans present some of the most significant challenges to sustainability, but they are also integral ecosystem components and must be considered as such in any effort to achieve sustainable- management goals. Given the growth in human populations, sustain-

A Framework for Sustainable Forest Management 795 able provision of ecosystem goods and services becomes an even more compelling goal. Humans as ecosystem components should be consiclered in at least three ways: I) their use of all resources in the ecosystem, 2) their use of the land, and 3) their effects as a result of nonindustrial private forests owners' decisions and goals. The principles outlined earlier in this chapter must form the founda- tion for forest management if the goals and challenges posed in this chapter are to be met over the long term. Making Management Adaptable Forest-management practices have long been adapted in response to changes ~ markets, social values, uncertainty and risk, and available information about the effects of management practices. Perhaps because change in these conditions in the Pacific Northwest is occurring at a more rapid pace than previously, or perhaps because the concepts of spatial and temporal scale of management are expanding, the term "adaptive management" is now being used in the region to describe a management scheme that pays particular attention to uncertainty and the general paucity of information about the effects of management on the sustainability of forests. Of necessity, natural resource managers set goals and implement practices with an incomplete knowledge base. Given the complexity and variability inherent in the natural world, this will always be the case. Thus, many treatments of ecosystem management have empha- sized that management should be viewed as experimental and that good management should include the means to learn from our experiments and adjust goals and practices accordingly. This concept of "manage- ment as experiment" is the basis for adaptive management. Waiters (1986) proposed the following elements for adaptive management: · Management practices should be framed by clearly stated objec- tives and bounded by an honest appraisal of the practical constraints on action. · Existing understanding of managed systems should tee represented in terms of explicit models of dynamic behavior that spell out assump- tions and predictions clearly enough so that errors can be detected and used as a basis for further learning.

796 Pacific Northwest Forests · Uncertainty and its propagation through time in relation to management actions should be understood using statistical measures and imaginative identification of alternative hypotheses (models) that are consistent with experience but might point toward opportunities for improved management success. · Policies should be balanced to provide for continuing resource production while simultaneously probing for better understanding and untested opportunity. Thus, adaptive management explicitly provides a basis for {earning as forest management proceeds to meet human-defined goals. it is based on a learning process that distinguishes among what is known, what is suspected, and what is not known. Adaptive management can be viewed as a series of actions that are tentative, that are watched closely, and that are changed as their results become apparent. It should be viewed as complementing, rather than replacing, reliance on formal research results. Monitoring the results of forest-management actions is needed to identify successful practices and to recognize when changes are necessary. Given the limitations in understanding of the behavior of, for example, managed forest ecosystems, managing without monitoring elements critical to meeting management goals is akin to trying to navigate without a compass (Lee 1993~. This is as true for monitoring social and economic factors as it is for monitoring ecosystem attributes. The scale and intensity of monitoring programs must be consistent with the scales of the processes that are to be monitored (Kevin 1992~. The Forest Service monitors some attributes of forest conditions at fairly gross geographic and temporal scales in the Forest Inventory and Analysis program (FlA) of its research branch. Although that level of monitoring provides some useful information on changes in forest conditions over periods of decades and at a multicounty spatial scale, it has not been designed to monitor changes in forest conditions at the level of a forest stand or in response to particular management actions. The scale at which attributes of a managed forest should be monitored needs to be addressed (Levin 1992~. Sampling design, technical aspects (e.g., logistics, costs, and equipment), and institutional and policy issues involving responsibility and coordination all need to be considered In monitoring to support effective adaptive management. Because forest ecosystems can buffer environmental changes that can be deleterious

A Framework for Sustainable Forest Management 797 over time, monitoring to detect early changes present a special chal- lenge. For example, trees and forests can look healthy even after irreversible damage from insects or pathogens. A clear understanding of the processes that underlie forest vitality is critical to the detection of early and often subtle indicators. That is one aspect of monitoring for adaptive management that needs further attention. Resolving Conflicts Processes for resolving conflicts in the management of Pacific Northwest forests clearly have not worked wed. Conflicts are caused by differences in value systems (e.g., the marketplace versus the political system), requirements of the present versus those of future generations, spatial issues (e.g., issues that cut across ownerships or political boundaries), and the time required for societal institutions (e.g., the legal structure, markets, and agencies) to respond and change. Yaffee (1994) suggests the need for four improvements in institutions to avoid or manage problems of the type exemplified in the disputes over the northern spotted owl: (~) new mechanisms to bridge the agency-nonagency boundary to build understanding and political concurrence; (2) altered approaches to organizational management, including updated notions of leadership; (3) improved means of gathering and analyzing information about resource problems, organizational possibilities, and political and social context; and (4) ways to promote a culture of creativity and risk-taking to generate more effective options for the future (Yaffee 1994~. It is evident from this listing that no single approach or new policy institution will solve the problems faced in managing the forests of the Pacific Northwest. In the spirit of adaptive management, various efforts in the region are being made to find ways to resolve some of the conflicts over forest management. Some of these are led by stakeholders such as the states, large landowners, or local groups of citizens. For example, under a process fostered by the Washington Forest Practices Board and enforced by the Washington Department of Natural Resources, landowners are cooperating in a program of watershed analyses to resolve issues involving timber, fish, and wildlife at the individual watershed level. Cooperative projects, some of which have been identified for "adaptive management units" under the Northwest Forest Plan, have been

798 Pacific Northwest Forests initiated by groups of citizens and governmental agencies in several parts of the region. Many articles and reports discuss such cooperative projects, including Wondolleck (1988) and Endicott (1993~. The FEMAT process for federal lands is another approach, one that has been driven by the need to accommodate forest management of federal lands to the requirements of the Endangered Species Act (ESA). Various "habitat conservation plans" (HCPs) have also been developed by owners of private forests in response to the requirement of the ESA that habitat for endangered species cannot be reduced if it will endanger protected wildlife. The specific objective of most of these HCPs is to avoid the "incidental take" provisions of the ESA regulations and thereby to reduce some of the uncertainty in dealing with these regulations. Other approaches include one proposed by Gordon (1994) that an effort be launched to extract a national public vision to underpin and guide forest policy and management. No single approach is best or, by itself, sufficient. The obvious need is to keep experimenting with ways to resolve conflicts in forest management, especially at the locale level, in full recognition of the existing limitations of the overall institutional structure for doing so.

Next: 9 Conclusions and Recommendations »
Environmental Issues in Pacific Northwest Forest Management Get This Book
×
Buy Paperback | $50.00 Buy Ebook | $39.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

People are demanding more of the goods, services, and amenities provided by the forests of the Pacific Northwest, but the finiteness of the supply has become clear. This issue involves complex questions of biology, economics, social values, community life, and federal intervention.

Forests of the Pacific Northwest explains that economic and aesthetic benefits can be sustained through new approaches to management, proposes general goals for forest management, and discusses strategies for achieving them. Recommendations address restoration of damaged areas, management for multiple uses, dispute resolution, and federal authority.

The volume explores the market role of Pacific Northwest wood products and looks at the implications if other regions should be expected to make up for reduced timber harvests.

The book also reviews the health of the forested ecosystems of the region, evaluating the effects of past forest use patterns and management practices. It discusses the biological importance, social significance, and management of old-growth as well as late-succession forests.

This volume will be of interest to public officials, policymakers, the forest products industry, environmental advocates, researchers, and concerned residents.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!