National Academies Press: OpenBook

Forest Health and Biotechnology: Possibilities and Considerations (2019)

Chapter: 4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees

« Previous: 3 Mitigating Threats to Forest Health
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

4

Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees

Any intervention to address forest health involves consideration of associated ecological, economic, social, and ethical issues. This chapter discusses these considerations related to using biotechnology to mitigate forest health threats. Some of these considerations are unique to biotechnology, but others are applicable to any intervention.

ECOLOGICAL CONSIDERATIONS

From an ecological standpoint, the use of biotechnology to improve forest health is fundamentally different from a decision to employ biotechnology for pharmaceutical or other existing human uses, including (to some degree) crops. A general principle for these other biotechnological products is that the intent is for them to remain confined to the production site. Pharmaceutical products can generally be produced in a laboratory or industrial setting and thus kept secure from the broader environment. Most crops (biotech and nonbiotech) have been bred to grow in mono-cultures that are planted and harvested on an annual basis. Although gene flow from crops is possible, escapes and hybridization with wild relatives are generally low (NASEM, 2016b). Conversely, in cases where biotechnological approaches are implemented to address forest health, the intention is to maximize spread of the modified genome into forests to confer increased genetic resistance to insect pests or pathogens throughout the range of the tree species. Furthermore, forest trees are perennials that grow and interact with many other species throughout their long life span.

The 2016 National Academies report on gene drives (NASEM, 2016a) identified several interacting factors that influence the success of a gene drive propagating in the environment, which parallels the spread of biotech trees for forest health (NASEM, 2016a:3)1:

___________________

1 The committee has replaced “gene drive” with “genetic modification.”

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
  • The evolutionary “fitness” of individuals carrying the [genetic modification]—that is, their ability to produce fertile offspring—as compared to individuals not carrying the [genetic modification].
  • The “conversion rate,” which describes how the [genetic modification] is passed to subsequent generations when one parent carries the [genetic modification] and the other does not.
  • “Gene flow,” which describes how the [genetic modification] moves between different populations of the target species.
  • “Horizontal gene transfer,” or the potential for [genetic modifications] to move from the target species into entirely different species.

For trees, the committee considers it important to add additional items to this list:

  • Establishment, which describes the ability of individuals carrying the genetic modification to compete with other individuals, allowing establishment and growth to reproductive maturity.
  • Standing genetic variation, which is the presence of alternative forms of a gene in a population.

Furthermore, the report on gene drives noted that ecological factors at the community level are important to consider. These include a species’ role in its community. Another factor to consider is the ability of a change to the species to create a tipping point from one ecological community configuration to another configuration. Genetic changes introduced into trees to address forest health threats have the potential to take on characteristics of invasive species that tip the balance of ecosystems.

Fitness and Conversion Rate

Attention to fitness and conversion rate will be critical because the intent of biotech trees would be to recover species over both large temporal and spatial scales (Newhouse et al., 2014). Even substantial outplanting efforts will provide only founding individuals intended to result in populations with increased resistance to the insect pest or pathogen involved. These trees will have to retain “fitness” to survive and “convert” that fitness, that is, produce viable propagules with the resistance in future generations. To ensure genetic fitness over the long term, the possibility of trade-offs between genetic traits needs to be examined, that is, whether modifying plants for resistance results in trade-offs in growth, drought resistance, or seed production (Lovett, 2018). Given that forest trees will be on the landscape for decades to centuries, the conversation rate of any genetic resistance will have to provide durable and heritable resistance under unmanaged or minimally managed conditions over time (Sniezko and Koch, 2017). Certain types of resistance may have limited or no utility if they cannot be sustained in the population (Kinloch et al., 2004; Sniezko and Koch, 2017).

Gene Flow and Horizontal Gene Transfer

Many of the tree species under consideration are wind-pollinated, suggesting the potential for long-distance gene flow within the target species (Liepelt et al., 2002; Van Deynze et al., 2016; Semizer-Cuming et al., 2017). Furthermore, the possibility of long-distance pollen dispersal has been considered a potential evolutionary mechanism for tree populations to withstand the negative effects of climate change (Kremer et al., 2012). Thus, the confinement of gene flow (i.e., based on

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

jurisdictional or cultural boundaries) would not be possible or even desired because presumably the trees with resistance will be the progenitors of future generations of the species. Additionally, interspecies gene flow, via horizontal gene transfer or hybridization, could also occur. Although the extent of horizontal gene transfer in plants is not well understood (Richardson and Palmer, 2007), several mechanisms of gene transfer between plant species have been described. Direct transfers can occur via parasitism, symbiosis, pathogens, epiphytes, entophyte, and grafting, and indirect mechanisms of horizontal gene transfer include pollen, fungi, bacteria, viruses, viroids, plasmids, transposons, and insects (see Gao et al., 2014, for a review). Even if the extent of horizontal gene transfer in plants is not well defined, hybridization between related plant species is common (Arnold, 1992). In the case of an introduced biotech tree, if hybridization with other species occurs, constraining such hybridization would be impossible unless hybrids had significantly reduced fitness (e.g., Ellstrand, 1992; Feurtey et al., 2017). Given that the tree species under consideration for this examination of biotechnology use on forest health are native species and that the introduced gene will likely spread within the native community, potential impacts both to the species involved and to the associated ecological and human communities need careful analysis (see section “Impact Assessment Framework” in Chapter 5; NASEM, 2016a).

Tree Establishment

For biotech trees to address forest health threats, they have to be competitive with their conspecifics and with other plant species in the ecosystem. Even if a biotech tree is genetically fit and able to convert its resistance to subsequent generations, it will not become established in a forest if it is not competitive in the ecosystem. This competitiveness needs to be balanced with any potential for the biotech tree to become a nuisance species, analogous to a nonnative invasive species that alters an ecosystem. Some characteristics to consider when it comes to establishment are growth rate, maturation age, fecundity, root suckering, understory establishment, and allelopathy (Clark, 2018).

Role of Trees in Ecological Communities

By virtue of their woody growth, trees are able to develop perennial tissue that provides the structure for developing height and access to sunlight above other plants in the forest, making flowers more visible and accessible for pollination, generally via pollinators or wind, and dispersing seeds over long distances via wind, water, or animals. Woody stems support both dead and living tissues, providing important resources for multiple species across seasons and decades to centuries. As a result, forests harbor substantial biodiversity (Brockerhoff et al., 2017). Additionally, tree longevity means that biogeochemical cycles are locally influenced by trees, which stabilize soils and can alter local and regional climate (Bonan, 1999). Modeled impact of tree loss on carbon and nitrogen cycling suggests that some impacts may last for centuries (Crowley et al., 2016).

Disturbances of multiple scales in forests maintain successional gradients and biodiversity, recycle nutrients, and control population dynamics. Regional and local environmental variation have similar impacts. While pests have always had a natural role in individual, stand, and landscape dynamics of trees, the scale of that role has been substantially expanded with climate change, forest management, and pest introductions (see Chapter 2).

Exposure of trees to pest species over many generations has resulted in development of resistance to impacts of both specific and generalized pests. The distribution of this resistance may not be consistent across the range of a tree species (see section below “Standing Genetic Variation in the Context of Range Position”), and resistance may involve trade-offs with other traits such as growth, drought resistance, seed production, tissue palatability, and nutrient dynamics that have implications for ecosystem function (Reid et al., 2016; Lovett, 2018). Additionally, the longevity of trees relative to their pests means that the latter have the potential to evolve more rapidly than their hosts.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

The Importance of Maintaining Standing Genetic Variation for Forest Health

An important difference in the forest tree situation from agricultural uses of biotechnology is that a focus on recovering forest species requires incorporating the specific genetic change while retaining the breadth of genetic diversity in forest populations. This diversity permits the species to continue to evolve under changing abiotic and biotic conditions (see “Site Management Practices” in Chapter 3). Provenance studies in many tree species have revealed substantial variation in response to environmental factors; variability that will be critical for potential adaptation of species to environmental changes (e.g., Aitken and Bemmels, 2016; Montwé et al., 2018). Adaptation to local environmental conditions often requires different breeding populations for different portions of the species’ geographic range. As a result, to maximize forest health, the genetic changes need to be incorporated into a diverse breeding population rather than into an individual cultivar or variety.

Standing Genetic Variation

Long-lived forest tree species often have wide geographical distribution and exhibit predominantly outcrossing mating systems. As a consequence, much of the genetic variation in populations of forest trees is partitioned within populations and very little among them (Hamrick, 2004) with some exceptions (see Kinloch et al., 2003). Standing genetic variation is the amount of allelic variation at a genetic locus that is segregating within a population (Orr, 2005). This variation, shaped by evolutionary and demographic forces, provides raw material for rapidly adapting to the changing environment as well as to novel habitats during range expansion (Barrett and Schluter, 2008). When an allele confers a functional benefit (e.g., resistance to an insect pest or disease), natural selection may act on it and drive it to fixation in the population. The genetic locus then no longer constitutes part of the standing genetic variation. Additionally, mutations arising de novo in populations may also provide a secondary source of genetic variation for adaptation through the action of natural selection, but the mutation rate is very low in forest tree species (Savolainen and Pyhäjärvi, 2007). Thus, standing genetic variation is the primary vehicle for evolutionary change and is highly consequential for forest health, the forest’s adaptability to environmental change, and the resilience of the forest to insect pests and pathogens.

Standing Genetic Variation in the Context of Range Position

The distribution of a species’ standing genetic variation across its range has historically been studied in terms of population position at range center versus range margin vis-à-vis the classical central marginal hypothesis (Eckert et al., 2008). Theoretical predictions of the hypothesis suggest diminished levels of genetic variation at range margins where environmental conditions are usually at the limits of physiological tolerance, in contrast with the abundant center where conditions are conducive to maintenance of optimal levels of genetic variation. While there is some support in literature for this prediction, the debate around this issue is not fully resolved. A synthesis of the evolutionary and demographic dynamics at various parts of a species’ distribution range (Hampe and Petit, 2005) may suggest that the warmer rear-edge populations may contain genetic variants preadapted to the environmental conditions that are forecasted for the northern latitudes under climate change. Experimental evidence is limited but is continuing to accumulate for this prediction (e.g., Rice et al., 1993). In particular, understanding the patterns of radiation out of the glacial refugia (i.e., geographic regions where flora and fauna survived during the ice ages and later recolonized postglacial habitats) and how that has shaped the standing genetic variation in response to past climates is important when choosing genetic backgrounds against which to deploy biotechnological solutions to climate or pest mitigation. Assisted migration by humans of forest trees to mitigate the

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

effect of climate change on tree species is being considered. In these cases, such preadapted variants may provide the key to healthy forests in geographically distant but environmentally similar (as a result of climate change) habitats. This knowledge base of population-level adaptive standing genetic variation together with data from ecological factors and climate modeling could provide clues to adaptability of forest tree populations to climate change through either migration or adaptation (Neale and Kramer, 2011).

Range expansion and contraction have been part of the evolutionary history of most tree species. During repeated climatic oscillations, advancing glaciation forced forest tree populations into refugia at both southerly latitudes in North America as well as in the Pacific Northwest (Shafer et al., 2010). Receding glaciers allowed species to expand to higher latitudes out of these refugia. Thus, the standing genetic variation in the refugia is a product of eons of demographic and evolutionary forces and thus likely to contain variants that are preadapted to a variety of environmental conditions. The current distribution of loblolly pine (Pinus taeda) is, for example, hypothesized to have radiated from two southern refugia, one in Florida and another in Mexico (Schmidtling, 2003). Similarly the glacial refugia for Populus are predicted to have existed in the Rocky Mountain region (Levsen et al., 2012), which constitutes the present-day southern-range edge of balsam poplar (Populus balsamifera), a boreal tree with one of the widest distributions in northern North America. Red maple (Acer rubrum) has been shown to have taken refuge in the unglaciated Appalachian Mountains, south of the Laurentide ice sheet (Delcourt and Delcourt, 1984). A similar pattern was observed in American beech (Fagus grandifolia) (McLachlan et al., 2005), whose refugium was located immediately south of the glaciation, whereas the western redcedar (Thuja plicata) is hypothesized to have expanded in the previously glaciated northern region out of a coastal refugium off Washington and south of the glaciation (Barnosky et al., 1987). Thus, historical distributions of tree species have contributed to their standing genetic variation, a factor that needs to be considered when assessing forest health and tree resistance to pests.

Local Adaptation and Its Genetic Basis

A genotype or a population is locally adapted when its fitness is higher under the local environment, but exhibits lower fitness elsewhere (Savolainen et al., 2007). In forest trees, many of these locally adaptive traits are complex in that their underlying architecture is controlled by multiple genes (Savolainen et al., 2007). Understanding the genetic basis of such complex traits remains a foremost goal in all of biology because of its implications for human health, agriculture, forestry, and ecosystem conservation and management. Given the postglacial phylogeographic history of forest tree species, local adaptation is likely widespread among their populations. Recent studies of many temperate forest tree populations have revealed strong latitudinal adaptation to the environment (e.g., temperature and photoperiod) for quantitative traits such as flowering phenology, growth, cold and drought tolerance, and ecophysiology (Howe et al., 2003; Savolainen et al., 2007; Aitken et al., 2008; Alberto et al., 2013; Olson et al., 2013; Guy, 2014). Landscape genetics/genomics is a relatively new field of research that aims to understand the landscape-level processes leading to local adaptation in widely distributed taxa (Manel et al., 2003; Manel and Holderegger, 2013). This field has leveraged the rapid advances and exponential growth in sequencing technology coupled with theoretical and methodological developments (Sork et al., 2013; Lotterhos and Whitlock, 2015) to facilitate genome scans to understand local adaptation.

Given their substantial diversity and complex phylogeography, the knowledge base of adaptive standing variation and local adaptation in forest trees is limited, but the availability of novel analytical tools (ecological, climatic, genomic, and computational) is paving the way for a better understanding of their adaptive potential. Considering the multitude of threats facing forests and projected climate change over the next century, some researchers have proposed assisted gene

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

flow between populations to facilitate forest survival in the 21st century and beyond (Aitken and Whitlock, 2013; Dumroese et al., 2015; Aitken and Bemmels, 2016). Any biotechnological approach, when combined with this assisted migration, would need to be fully informed by a thorough understanding of the extent of standing genetic variation and patterns of local adaptation within the species. One species where such studies have been conducted is whitebark pine, whose populations have sustained heavy losses due to a multitude of factors (see section “Case Study Trees” in Chapter 2). Rangewide and local-scale population genetic analyses and common garden experiments have revealed adaptation for growth and survival (Warwell, 2015), cold (Bower and Aitken, 2008), drought (Warwell and Shaw, 2017), soil water availability (Lind et al., 2017), and growth rhythm (Warwell and Shaw, 2018) along climatic gradients in whitebark pine. This finding has led to the development of seed transfer (assisted migration) guidelines for restoration purposes (Bower and Aitken, 2008), which have subsequently been employed to test the potential of the species to grow successfully beyond its northern range-margin (McLane and Aitken, 2012). The results from that research suggests that assisted migration could be a promising first step in the restoration of threatened species when information on standing variation and local adaptation is coupled with climate and species distribution modeling. A series of common garden field trials of whitebark pine have recently been established to help more fully understand the adaptive genetic variation in whitebark pine and provide field validation of resistance to white pine blister rust (Cartwright et al., 2016; Cartwright, 2018; USDA-FS, 2018). Monitoring these trials over the next several decades will improve understanding of how to best restore whitebark pine forests in the future.

Disruption of Local Adaptation Under Climate Change

When considering assisted migration as a potential restorative tool, it would be important to keep in mind that impending climate change may disrupt existing gene × environment associations (i.e., local adaptation), resulting in maladaptation. For widely distributed tree taxa, some parts of the range may be more vulnerable to such disruption than others. For example, strong adaptation to climate and photoperiod in phenological traits is known to occur in the case of Populus balsamifera (Soolanayakahally et al., 2009; Keller et al., 2012). Fitzpatrick and Keller (2015) demonstrate through modeling of various future climate scenarios where along the spatial landscape local adaptation will be disrupted. Using variation in GIGANTEA 5, a circadian-clock gene with strong local adaptation to temperature, they show that northernmost populations of P. balsamifera will likely experience the largest “genetic offset” from the adaptive optima. Genetic offset due to future patterns of changing rainfall have also been predicted in populations of Hawaiian koa trees (Acacia koa) and may inform reforestation and seed transfer guidelines (Gugger et al., 2018).

Thus, a biotechnological approach would need to account for the standing genetic variation and the extent of local adaptation and avoid swamping local adaptation or introducing maladapted genotypes while also supporting natural gene flow. It would also be prudent to identify populations that may likely experience genetic offset due to the disruption of local adaptation through climate change.

ECONOMIC CONSIDERATIONS

The committee identified the economic considerations of deploying a biotech tree resistant to insect pests or pathogens through the resulting impacts on ecosystem services. First, the significant economic value of forests to humans is one important motivation for intervening to maintain or restore forest health through the introduction of a biotech tree. Second, the incentive for the public and private sectors to invest in the development of such a biotech tree is conditioned by the types of benefits its introduction is expected to provide, largely dependent on whether the anticipated

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

gains in goods and services are traded and priced in markets. Third, the eventual introduction of the biotech tree will raise issues around consumer acceptance of the goods and services obtained, preferences that may be reflected in third-party certification schemes.

Economic Value of Forests

The natural resources of the U.S. forests support both private and public uses. In the United States, one-third of all land area is forested, more than 300 million hectares out of 980 million total (Oswalt et al., 2018). Timberland accounts for two-thirds of forestland, with about 13 percent of that land being planted and the rest in forest of natural origin. With regard to ownership of forestland, just over 40 percent nationally belongs to federal, state, tribal, local, and municipal governments, with the other 60 percent in private hands. Changes in nondisclosure laws pertaining to data-reporting mean that it is no longer possible to distinguish corporate from noncorporate private ownership. However, to the extent that corporate ownership is associated with tree plantations, the lower bound of corporate control could be around 9 percent of forestland, which would be about 13 percent of timberland. These percentages vary regionally. For example, public entities own three-quarters of timberland in the intermountain West, but only about one-fifth in the Southeast. Overall, U.S. forests vary with respect to ownership and geographic distribution (see Oswalt et al., 2018, for more detail).

There are three main components of the private forest industry (U.S. Census Bureau, 2017):

  1. Forestry and logging (growing and harvesting on a long production cycle, more than 10 years);
  2. Wood product manufacturing (lumber, plywood, veneer, containers, flooring, etc.); and
  3. Paper manufacturing (pulp, paper, paper products).

As components of gross domestic product (GDP), altogether these sectors accounted for about 0.5 percent in 2017 (BEA, 2018). Wood-products and paper-sector value added was about 5 percent of manufacturing GDP, which itself accounts for 12 percent of GDP. Employment in all three sectors has been just under 1 million people in recent years, out of a total workforce of 125 million (BLS, 2018).

The value added by activities on public lands is harder to quantify because of the nonmarket, noncommercial nature of ecosystem services. National income accounts do not recognize the contributions of forests in providing fish and wildlife habitat protection, watershed protection, carbon sequestration, and many other ecosystem services (FAO, 1998). Integrated economic and environmental accounting would provide a more complete picture of U.S. wealth and income from forests, but this approach has not been pursued by statistical agencies at the national level in recent years. The U.S. Bureau of Economic and Business Affairs produces satellite accounts for outdoor recreation (about 2 percent of GDP in 2016) and for travel and tourism (about 3 percent in 2016). Forests, of course, support only some part of these activities.

Outside the official national accounts, researchers have employed various techniques to value forests in monetary terms. Based on the concept of ecosystem services, one estimate pegged the value of boreal and temperate forests at $3,137 per hectare per year (2007 U.S. dollars) (Costanza et al., 2014). Of the global total, the United States is about 8 percent or 310,095,000 hectares (FAO, 2016). Based on that estimate of area and per-unit value, the total value of U.S. forest ecosystems would be roughly $975 billion (for comparison, U.S. GDP in 2007 was about $14 trillion). With a narrower focus, the recreational value of the U.S. National Forest System has been estimated at about $14 billion averaged over the period of fiscal years 2011 to 2015 (Rosenberger et al., 2017). Filtration of airborne particulates by U.S. forests in 2010 had an estimated value in human health benefits of $6.8 billion (Nowak et al., 2014). These estimates are made with varying assumptions

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

and can only be considered very approximate; valuation of ecosystem services is a fraught task, and some important services, particularly associated with nonuse value, are not easily expressed in monetary terms at all. However, such estimates do suggest robust value for the nation’s forests, even without taking forests’ noninstrumental values into account. In some cases, it may be possible to avoid diminution of that value due to damage by insects and diseases, and possibly even reverse losses, by the introduction of biotech trees. Still, assessment of the value to be restored or protected by the introduction of any one tree can only be made with reference to the particular uses and characteristics of the forest ecosystem in question.

Incentives to Invest in the Development of a Biotech Tree

When a forest ecosystem is threatened, the motivation to restore it to health is conditioned by concerns about the ecosystem services that are adversely affected or diminished when a tree species is lost or declining. The introduction of a biotech tree may thus be expected to affect the services provided by the forest ecosystem of which it is a part. Changes in these services can imply changes in the human uses and benefits derived from these services. To be able to assess these changes in terms of their economic significance, it is useful to review the classification of ecosystem services in a use/nonuse framework of total economic value (Pearce et al., 2006). Use value arises from the actual, planned, or possible use of a service. For example:

  • An extractive use might be the harvest of timber for use as a biofuel.
  • An in-situ use might be hiking or bird watching or contemplating the aesthetic beauty.
  • Option value might arise because of a desire to preserve the possibility of future use of the forest and its services.

Nonuse value stems from the benefit that arises even if there is no actual or planned use of the service.

  • Existence value might derive from the knowledge that old-growth forests exist, even if no human visit were ever to occur (see section “The Value of a Healthy Forest” in Chapter 2 for further discussion).
  • Bequest value would reflect a concern for the ability of succeeding generations to use the services, as with the 1892 creation of “forever wild” Adirondack Park.

Uses and nonuses affected by the introduction of a biotech tree may be embodied in market goods (such as the extractive use of timber for biofuel) or may be considered as public goods or benefits not traded or priced in markets (such as the existence of a stand of old-growth trees). These distinctions matter when evaluating incentives to develop biotech trees.

Trees, once planted and maturing, can provide both use and nonuse values as reflected in either public or private benefits. Public benefits are those that cannot be exclusively captured by an individual or a firm but are shared across many people and communities. Examples include clean air and water, flood control, support for biodiversity, and scenic landscapes (USDA-FS, 2007). In contrast, private benefits flow from extractive uses (such as logging), that is, products that can be exclusively held and sold in markets by firms and individuals. As would be expected, then, investment in the development of trees to provide mainly public goods—such as resistance to insect pests and pathogens—is usually pursued by governments and by nonprofit entities. Commercial forestry is the purview of firms with a profit motive.

The significance of the mix of public and private benefits of a biotech tree to the incentives to develop it can be illustrated with two examples. One is the freeze-tolerant eucalyptus for private

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

planted forests and the other is the blight-resistant chestnut tree for less managed public and private forests.

  • The freeze-tolerant eucalyptus is to extend northward toward the range of its use as a plantation tree in the southeastern United States. These nonnative, biotechnological trees are fast-growing and could provide timber and pulpwood, as well as a source of lignocellulose for the production of energy and advanced biofuels (Hinchee et al., 2011). Although these trees can provide public benefits as well (such as effects on air quality), the main incentive for their development is the potential for the sale of their products in markets. Accordingly, the private firm Arborgen has been the developer of the freeze-tolerant tree.
  • The blight-resistant chestnut tree similarly can provide benefits in the form of marketable commodities (e.g., wood and chestnuts), but its developers are organized as a university and a nonprofit foundation. The release of the tree into less managed or unmanaged forests may result in its restoration as a key ecosystem species and also in the aesthetic and cultural appreciation it enjoyed when it dominated regions of the northeastern U.S. forests (see section “Case Study Trees” in Chapter 2). These aims have strong public-good aspects.

Private firms have the incentive to introduce a tree that provides appreciable returns to sale of its products in markets, whereas public entities are more likely to value its ability to provide public goods that are unlikely to generate market revenue. It is the case that federal and state governments may earn some revenue from timber sales on public land and may support research to enhance productivity of commercial forestry. In the current context, however, public interest would be in innovations that protect forest health, broadly defined and with many public good aspects.

The potential for a biotech tree to yield market and/or public goods and services depends on the particulars of the changes in uses and nonuses that occur when it is introduced into a forest ecosystem. Depending on the mix and the size of the market or public benefits to be had, the private or the public sector may take the lead in development. Moreover, the incentive to apply biotechnology in trees is strongly conditioned by the relatively long time between a tree’s planting and its reaching maturity. Compared to agricultural field crops, which are typically harvested annually, tree crops’ life cycle covers multiple growing seasons, measured in years. As a result, the benefits of tree development and planting may not accrue until far into the future, whereas much of the cost of planting may occur right away. In such circumstances, the economics depend on the time value of money, that is, the opportunity cost of using funds to plant trees and wait for returns versus investing the money in an activity that yields returns much sooner.2 This element of delay between planting and maturity has significance for decision making in both the public and private sectors.

The aims of introduction of a biotech tree may be more oriented to generation of revenue from market or more aligned with outcomes associated with public goods. Whether it is a private or a public interest, resources are required to support the research and development (R&D) necessary to bring forth a biotech tree. The potential for forest biotechnology depends on public and private investment in its R&D. The 2002 National Research Council report National Capacity in Forestry Research found that, despite apparently large returns to forestry R&D, there were significant gaps in basic biological knowledge and deficiencies in understanding of forest health, systems, and management and wood science (NRC, 2002). More recently, a blue ribbon panel report from the U.S. Endowment for Forestry and Communities asserted that innovation in the sector has slowed over the past several decades (Jolley et al., 2017). This panel estimated U.S. forestry R&D at $700

___________________

2 See Sedjo (2006) and, for a detailed exposition of the use of the discount rate in decision making (especially for regulatory purposes), see EPA (2014).

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

million annually, with the federal government accounting for $500 million, state governments for $150 million to $175 million, and nongovernment entities for $10 million to $15 million. The panel did not address biotechnology specifically, but it did note opportunities to enhance both traditional forest products (e.g., engineered solid wood products and midrise buildings) and nontraditional forest products (e.g., carbon sequestration, renewable energy production).

The willingness of private firms to invest in research is a function of costs and expected returns. The panel from the U.S. Endowment for Forestry and Communities reported that corporate research in the forest sector, at 0.5 percent of annual revenues, is far below that of other sectors, such as biomedical science (including molecular biology) and health care (almost 12 percent of revenues) and automobiles (3.5 percent). The portion devoted to biotechnology was not estimated. Despite this relatively low level of investment, there are incentives to invest in insect pest and pathogen control and in product innovation. For example, losses to timber producers due to tree damage by the southern pine beetle (Dendroctonus frontalis) have been estimated at $43 million annually and $1.2 billion over about a 30-year period (Pye et al., 2011). Even if conventional tree breeding and biotech tree development had comparable costs for R&D, use of biotechnology comes with the added expense associated with regulatory approval.3 Compliance costs associated with the development of herbicide-resistant maize have been estimated at $6.0–$14.5 million and for insect-resistant maize at $7–$15 million (Kalaitzandonakes et al., 2007). If forest biotechnology compliance costs were comparable, two or three passes through the regulatory system would represent about 5 percent of all forestry R&D spending. Whether those costs can be justified will depend, of course, on the particulars of the tree under consideration, but they are a unique factor in assessing investment in biotechnology R&D prospects. Given the relatively long time to benefits accruing with tree maturity, increasing the upfront costs of development may adversely affect the incentive to invest in forest trees.

Public investment in forestry research may be intended to support private-sector economic development, or be aimed at enhancing the societal benefits forests that provide, or both. The federal government has been responsible for the largest investment in tree breeding in concert with states (Jolley et al., 2017). Nonprofit entities, such as foundations, may also invest, as has been the case with the American chestnut. Underinvestment in public goods is a market failure that can be addressed by public intervention. Motivating and justifying spending on forest R&D depends on the identification of the scope and magnitude of the public benefits arising from improvements in forest health and in ecosystem services. (This proposition is true for any kind of effort, whether it be targeted to the use of biotechnology or not.) Consequently, multidisciplinary efforts to characterize these benefits play a role in creating arguments for public expenditures (Boyd et al., 2016). It is also the case that these benefit estimates have value in the context of regulatory decision making when weighed against the costs and risks of introducing a biotech tree. Although the challenges to assessing nonmarket values of these public goods are considerable, there are nonetheless methods (such as contingent valuation) currently being employed in the regulatory process (EPA, 2014).

Proposals for forestry R&D compete with other demands on public resources. Given that the benefits of tree development may occur far into the future, it can be challenging to acquire public funding given competition from activities with a more immediate payoff. Moreover, because the beneficiaries of the investment may be numerous and geographically dispersed, advocacy by a distinct constituency may not be present. Public concern about the use of biotechnology in a forest setting (discussed in the section below, “Current Research on Public Views”) may also affect the willingness of public officials to commit resources. In fact, public spending on forestry R&D has not increased in recent years. There is recognition of this situation in calls for public–private

___________________

3 Costs associated with the regulatory approval process in the United States for biotech plants is discussed in detail in Chapter 6 of Genetically Engineered Crops: Experiences and Prospects (NASEM, 2016b).

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

partnerships to pursue projects that had in the past been funded by federal and state sources (Jolley et al., 2017).

Whether R&D is done in the public or private sector, the assignment of intellectual property rights associated with an innovation can affect its adoption and use. A private firm may seek to recoup the upfront costs of development and regulatory approval by patenting that protects its revenue stream and possibly enhances it through licensing. However, to the extent that the public sector would wish to enable the spread of an insect- or disease-resistant biotech tree, then patenting would not seem to be consistent with widespread adoption. Graff and Zilberman (2016) discuss this dynamic of the public interest in “orphan” crops not attractive for commercial purposes and suggest that the presence of significant social benefits motivates development by the public sector. In addition, ethical objections to patenting may arise, for example, over the propriety of patenting living organisms.

The emergence of CRISPR and other genome-editing techniques raises questions about their use by the public sector in development of a resistant biotech tree. At the time the committee was writing its report, there were a number of competing patents and patent applications (Cohen, 2018). However, one key patent holder, the Broad Institute, will not require licensing of its CRISPR innovations for academic and nonprofit use. Specifically, it states, “Nonprofit institutions and government agencies do not need to receive a written license from Broad to conduct internal research, including sponsored research, to the extent that such research does not include the production or manufacture for sale or offer for sale or performance of commercial services for a fee.”4 So it would seem that CRISPR-enabled resistance in a biotech tree would preclude the tree’s commercial sale, which might well be consistent with the public interest in protecting forest health.

Consumer Preferences

Another consideration specific to the development of biotech trees is the nature of the markets into which private goods such as timber and pulpwood are sold. Motivated by concerns about the sustainability of the use of forests globally, consumers have sought assurance that the tree products they buy are produced with practices compatible with that aim. Usually, it is not obvious from inspection of the product how the tree was grown. Consequently, third-party organizations now provide certification to consumers that the trees used for making the product are cultivated and harvested in accordance with specified methods that are supposed to promote sustainable use of forest resources.5 At the time the committee was writing its report, some forest certification programs applied in the United States prohibited the use of biotechnology. For example, based on its embrace of the precautionary principle,6 the Sustainable Forestry Initiative (SFI) has restrictions on the use of biotech trees until the end of 2022. However, SFI policy statements recognize the evolving nature of the underlying science and the potential benefits of biotech trees and state that it will “proactively review and update … this policy as necessary” (SFI, 2015). Depending on the extent of adoption of such certification in the market, the introduction of biotech trees may be discouraged. At the time the committee wrote its report, it was difficult to judge whether such prohibitions against the use of biotechnology will endure as knowledge about the science and the specific properties and expected impacts of biotech trees become better known.

___________________

4 Broad Institute. Information about licensing CRISPR genome editing systems. Available at https://www.broadinstitute.org/partnerships/office-strategic-alliances-and-partnering/information-about-licensing-crispr-genome-edi. Accessed November 21, 2018.

5 FAO. Sustainable Forest Management (SFM) Toolbox. Available at http://www.fao.org/sustainable-forest-management/toolbox/en. Accessed November 21, 2018.

6 According to the European Parliament, “The precautionary principle enables decisionmakers to adopt precautionary measures when scientific evidence about an environmental or human health hazard is uncertain and the stakes are high” (www.europarl.europa.eu).

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

SOCIAL AND ETHICAL CONSIDERATIONS

Alongside ecological and economic considerations, any proposed use of biotech trees involves taking a variety of social and ethical issues into account. Social science studies provide some information about public views, and how the use of biotechnology in forests is likely to be understood by the public in relation to risk and to alternative tools for addressing threatening forest diseases. Other concerns relate to the ways people value forests, including forests’ wildness and naturalness, how people interpret the use of biotechnology as a conservation tool, and how the use of biotechnology in forests is likely to affect social justice.

Current Research on Public Views

A number of studies have examined societal views on forest health threats such as insect pests and pathogens (e.g., Flint, 2006; McFarlane and Watson, 2008; Chang et al., 2009; Flint et al., 2009; Müller and Job, 2009; Mackenzie and Larson, 2010; Kooistra and Hall, 2014; Poudyal et al., 2016). However, in comparison to the many studies examining societal views about the use of biotechnology in agriculture and food (e.g., Finucane and Holup, 2005; Costa-Font et al., 2008; Frewer et al., 2013; Lucht, 2015), there have been substantially fewer on views toward using biotechnology in trees to address forest health threats (for a table of these studies, see Appendix D). Most of these studies have been conducted in Europe and Canada, with only a few in the United States. The primary focus of most of these studies has been on biotechnology use aimed at increasing timber in forest plantations (e.g., faster tree growth, better wood structure) or responding to climate change, with only a few recent studies on using biotechnology for restoring tree species or reducing insect pests and pathogens in forests.

Understanding societal responses is important for multiple reasons. First, democratic governance of emerging technologies in forests and elsewhere requires attention to societal views and concerns. Second, given the variety of approaches available when confronting a threat to forest health, scientists and land managers may wish to align their actions with the preferences and values of citizens and other groups to build confidence and trust, avoid controversy, or both. Third, since the advent of the environmental movement in the United States in the 1960s, a broad spectrum of people demand and expect involvement in natural resource management issues with some even seeking co-management with agencies. Complicating these objectives, however, is the existence of multiple “publics” with interests in trees and forests (e.g., citizens or residents, government agencies, companies such as growers and processors, direct and indirect consumers of forest products, and environmental groups; Davison et al., 1997; Sedjo, 2010). Discontented groups can resort to administrative appeals, court cases, protests and demonstrations, ballot initiatives, and direct action if they perceive that their concerns are not being addressed. These societal responses, which can be influenced by the attitudes, norms, perceived risks and benefits, knowledge, trust, and values among citizens and other publics (e.g., agencies, companies), may affect the adoption and management of biotechnology tools in trees and forests (Sedjo, 2006; Gupta et al., 2012; Strauss et al., 2017).

Attitudes and Norms

Although the potential use of biotechnology in trees and forests has raised concerns among some people, as illustrated by monikers and catchphrases such as “Frankenstein Forests,” “Frankentrees,” “Designer Trees,” “Silent Forests,” and “Terminator Genes” (Hall, 2007; Gamborg and Sandøe, 2010; Lombardo, 2014; Porth and El Kassaby, 2014), a majority of the relatively limited number of published studies have shown somewhat positive attitudes and normative acceptance among the majority of citizens and several other publics (e.g., agencies, companies). Attitudes involve evaluating something, such as biotechnology, with some degree of favor or disfavor (i.e.,

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

like, dislike; good, bad), whereas norms are standards that individuals use for evaluating their acceptance of something and whether they think it should be allowed (Eagly and Chaiken, 1993; Vaske and Whittaker, 2004; Fishbein and Ajzen, 2010).

Most studies have reported that although people are most supportive of natural regeneration, selective breeding and planting of native tree species, and site management practices such as thinning and felling to address forest health threats, the majority also support some biotechnological approaches, and many of these methods are viewed as more acceptable than doing nothing in the face of severe threats to forests (Hajjar et al., 2014; Hajjar and Kozak, 2015; Nonić et al., 2015; Fuller et al., 2016; Needham et al., 2016; Jepson and Arakelyan, 2017a,b). Both Hajjar et al. (2014) and Hajjar and Kozak (2015), for example, reported that approximately 85–90 percent of their respondents living in Western Canada accepted the breeding of native species to address forest health threats from climate change, and approximately 50 percent also accepted the planting of trees with traits introduced via biotechnology, but only 35 percent accepted no interventions. A more recent study also in Western Canada, however, found that only 25 percent of residents were supportive of using biotechnology in reforestation efforts in response to climate change (Peterson St-Laurent et al., 2018). Needham et al. (2016) found that U.S. residents considered native tree breeding and other conventional forms of forest management to be most acceptable for addressing chestnut blight in American chestnut trees (68–88 percent), but a majority also supported using various types of biotechnologies for mitigating this issue (53–64 percent). Nonić et al. (2015) found that 56–59 percent of students in universities in Serbia agreed with using biotechnology in trees, and Fuller et al. (2016) reported that 66 percent of UK residents accepted biological control methods (including biotechnology approaches) for managing tree pests and diseases. Adding additional nuance, Jepson and Arakelyan (2017a) found that cisgenic approaches (i.e., genetic modification with genes largely from the same species) were in the top three preferred options (for 54 percent of residents surveyed) among eight courses of action for addressing ash dieback in the United Kingdom, but transgenic methods (i.e., genes from unrelated species) were the least or second-least preferred choice for 85 percent of these residents, suggesting that there is sensitivity to differences among various biotechnological solutions.

Many studies have also reported that biotechnological tools tend to be either just as acceptable or even more acceptable than hybridization with nonnative species (a nonbiotechnological tool). Needham et al. (2016), for example, reported that changing genes in American chestnut trees (57–58 percent of U.S. residents supported this approach in general), such as adding genes from bread wheat (the oxalate oxidase [OxO] gene; 54–55 percent supported this approach in particular), was more acceptable for addressing chestnut blight than breeding with nonnative Asian chestnut species (43–46 percent supported). Jepson and Arakelyan (2017b) found that only 17–18 percent of their respondents in the United Kingdom preferred breeding and planting nonnative ash to mitigate ash dieback, whereas 27 percent preferred “using genetic modification techniques, including cisgenics and transgenics” (unlike Jepson and Arakelyan, 2017a, cisgenic and transgenic approaches were combined in this later study by the same authors). Both Hajjar et al. (2014) and Hajjar and Kozak (2015) found similar levels of acceptance (approximately 50–60 percent) among Western Canadians for breeding and planting different tree species and planting species altered using a biotechnology approach for addressing threats to forest health from climate change. A more recent study in Western Canada found that 26 percent of residents supported reforestation with nonnative species and, similarly, 25 percent supported using trees containing genetic material altered through biotechnology (Peterson St-Laurent et al., 2018).

Surveys have found that the use of biotechnology in trees and forests is slightly less acceptable than the use of biotechnology in some other industries such as medicine, but more acceptable than using biotechnology in agriculture and food (Connor and Siegrist, 2010). Studies have also shown that people tend to be more supportive of using biotechnology to address immediately pressing and

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

tangible forest health threats, such as insect pests and pathogens, compared to other issues such as climate change or improving tree growth and productivity for increased timber harvesting (Nonić et al., 2015; Fuller et al., 2016; Needham et al., 2016). In a study of university students in Serbia, Nonić et al. (2015) found that, on average, enhancing resistance to diseases was the most acceptable use of biotechnology in trees. Similarly, Needham et al. (2016) reported that biotechnological approaches were slightly more acceptable among American residents for addressing chestnut blight (53–64 percent) than for mitigating effects of climate change (45–58 percent) or for increasing forest growth and productivity (43–55 percent). Despite these findings, the use of biotechnology is much more acceptable in plantation forests than in noncommercial forests. Jepson and Arakelyan (2017a), for example, found that only 38 percent of UK residents approved of planting cisgenic or transgenic ash trees in woodlands, but 60 percent supported planting these trees in forest plantations. Likewise, Kazana et al. (2015, 2016) reported that the majority (56–93 percent) of university students in 15 European and non-European (e.g., Argentina, Israel) countries approved of growing transgenic trees in plantations.

Attitudes and norms toward the use of biotechnology in trees and forests also vary among citizens and other interest groups (e.g., agencies, nongovernmental organizations [NGOs], scientists; Friedman and Foster, 1997; Strauss et al., 2009; Needham et al., 2016; Nilausen et al., 2016). Nilausen et al. (2016), for example, found that government (78 percent support) and industry (100 percent) representatives had highly positive attitudes toward using tools related to biotechnology in trees and supported their continued research and use in Canada, whereas environmental NGOs (50 percent) and indigenous groups (17 percent) had far less positive attitudes. Needham et al. (2016) reported that, compared to American citizens (53–64 percent support), other groups involved in forest issues (e.g., scientists, agencies, companies) had more positive attitudes (81 percent) toward using biotechnology to address chestnut blight. Conversely, Hajjar et al. (2014) found that although approximately 50 percent of Western Canadian residents supported planting genetically engineered trees to help forests respond to climate change, only 30 percent of community leaders (e.g., elected officials such as mayors, council members, and town managers) supported this approach. Friedman and Foster (1997) reported that government managers and scientists in the United States were concerned about potential impacts of using biotechnology in trees and forests (e.g., possibility of reducing genetic diversity). However, in a study of government, university, and private-sector scientists in both Canada and the United States, Strauss et al. (2009) found that greater than 70 percent believed regulatory requirements are significant impediments to research on forest biotechnology, and that the use of biotechnology in trees should be encouraged.

Perceived Risks and Benefits

These attitudes and norms associated with the use of biotechnology in trees and forests are influenced by cognitive factors such as risk perceptions (Connor and Siegrist, 2010), which are subjective and negative evaluations of threats posed by potential hazards (Slovic, 2000, 2010). Objective risk is defined as the calculated probability and consequences of potential hazards, whereas perceived risks are subjective judgments that draw upon intuitions and other heuristic processes (Slovic, 2000, 2010). For example, people often express more concerns about new, unknown, and unnatural hazards that are not well understood or are outside of their control (Slovic, 2000, 2010; Sjöberg, 2004; Finucane and Holup, 2005), and this means that these risks are often perceived as more significant than other hazards that are more common and well known, but have higher objective risk.

Given their novelty, biotechnological approaches for modifying forest trees are expected to be perceived as riskier than familiar methods such as selective breeding (Strauss et al., 2017). As mentioned above, research examining societal responses toward methods for mitigating forest

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

health threats showed that biotechnological approaches were often perceived as less acceptable than more familiar approaches such as tree breeding, planting, thinning, and felling (Hajjar et al., 2014; Hajjar and Kozak, 2015; Needham et al., 2016; Jepson and Arakelyan, 2017a,b; Peterson St-Laurent et al., 2018). Yet, such responses are not entirely consistent, because breeding with related species (e.g., American chestnut with nonnative Asian chestnut, native ash in the United Kingdom with nonnative ash) was considered by citizens to be riskier compared to some cisgenic (e.g., altering genes in native ash in the United Kingdom; Jepson and Arakelyan, 2017a,b) and transgenic approaches (e.g., inserting genes from unrelated species such as from bread wheat [i.e., the OxO gene] into the American chestnut; Needham et al., 2016), suggesting that perceptions of naturalness or familiarity may elicit different concerns in some cases. For example, Asian chestnut species are more closely related to the American chestnut than wheat is, but are not as familiar to Americans as wheat (i.e., as a source of bread; Strauss et al., 2017). Although speculative, it may also be that concerns about naturalness and purity of species are driving some preferences, as the transgenic tree maintains a higher percentage of American chestnut DNA than a backcrossed tree with DNA from both the American and Asian chestnut species (Nelson et al., 2014; Powell, 2014). In fact, recent research in Western Canada showed that perceived transgressions of naturalness drive resident perceptions of risk and levels of support more than the biotechnological intervention itself (Peterson St-Laurent et al., 2018).

Sjöberg (2004) identified interfering with nature and severity of consequences as two other important dimensions of risk perceptions related to biotechnology in general (not necessarily related to forestry or in any specific location). The same societal concerns about risks associated with humans manipulating, tampering, and interfering with nature have been found in studies of biotechnology in trees and forests in Western Canada (Hajjar and Kozak, 2015) and the United Kingdom (Jepson and Arakelyan, 2017b). People have also perceived that severe changes in ecosystem components and functions are among the largest risks of forest biotechnology. For example, respondents in several studies in Canada and Europe were concerned that altering genes or adding transgenes in plantation trees could change genes or reduce the genetic diversity of wild or native trees (through gene flow), causing long-term impacts on biodiversity that are currently unknown (Kazana et al., 2015, 2016; Nonić et al., 2015; Tsourgiannis et al., 2015, 2016; Nilausen et al., 2016) and potentially increasing legal and liability concerns (Strauss et al., 2009). Additional concerns included vulnerability to other tree diseases and impacts from more pesticide inputs (i.e., from using disease-, herbicide-, and insect-resistant traits; Kazana et al., 2015, 2016; Nonić et al., 2015). Taken together, these findings are consistent with those of Lorentz and Minogue (2015), who examined perceived risks of eucalyptus plantations in the southeastern United States and found that invasion potential and associated negative ecological impacts on nonplantation forests were primary concerns. Similarly, Friedman and Foster (1997) surveyed U.S. forest agency employees7 and found that loss of adaptation, reduced genetic diversity, and changes in ecosystem components were the largest perceived risks from the use of biotechnology in trees and forests.

In addition to these risks, attitudes toward using biotechnology in trees and forests may also be related to the extent that people view these approaches as beneficial. Research mostly in Canada and Europe has found that respondents in several studies perceived benefits of forest biotechnology, including

  1. Increased tree growth and productivity (Kazana et al., 2015, 2016; Nonić et al., 2015);
  2. Economic and community benefits such as greater employment and income, more economic diversification and competition, and reduced production costs and losses (Neumann

___________________

7Friedman and Foster (1997) surveyed employees of the U.S. Forest Service, the Fish & Wildlife Service, and the Bureau of Land Management as well as those at state and local agencies.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
  1. et al., 2007; Tsourgiannis et al., 2013, 2015, 2016; Hajjar et al., 2014; Hajjar and Kozak, 2015);
  2. Greater consumer choice and purchasing options (Tsourgiannis et al., 2013, 2015, 2016);
  3. Restoration of contaminated soils (Kazana et al., 2015, 2016);
  4. Reduced pressure on harvesting trees from forests if biotechnology is constrained to plantations focused on increasing growth (Nilausen et al., 2016); and
  5. Reductions in insecticide, pesticide, and herbicide inputs (Kazana et al., 2015, 2016).

These benefits, however, may differ depending on the type and scale of production (e.g., plantation owners or smaller community forests) as well as the intent of biotechnology use (e.g., to increase timber, protect or restore forests, reduce insects and diseases; Strauss et al., 2017).

Knowledge, Trust, Values, and Communication

Knowledge can also influence attitudes toward biotechnology (Connor and Siegrist, 2010). Some studies in Europe and the United States have shown that the majority of citizens and other publics (e.g., agencies, companies) who have been surveyed are aware of particular forest health threats and have heard about the potential for using biotechnology in trees and forests (Kazana et al., 2015, 2016; Nonić et al., 2015; Needham et al., 2016). Both Kazana et al. (2015, 2016) and Nonić et al. (2015), for example, found that 60–70 percent of their respondents in mostly European countries knew what a transgenic tree was and were aware of the meaning of transgenic trees. However, given the complexity and novelty of biotechnology in trees and forests, many people lack detailed knowledge about specific aspects of this topic (Strauss et al., 2017). Kazana et al. (2015, 2016), for example, found that despite high general awareness about this issue, fewer than half of their respondents in 15 European and a few non-European countries (Argentina, Australia, Israel) could specify benefits or risks and knew whether these trees could be deployed in plantation forestry (e.g., grown commercially, sold on the market). This finding led these authors to believe there was “a serious perceived lack of knowledge about potential benefits and risks of the cultivation of transgenic forest trees” (Kazana et al., 2015:344). Although this does not invalidate the findings of other social science research that can be valuable even when knowledge about specific aspects of a topic is relatively low, it serves as a reminder that as various publics learn more about biotechnology in trees and increase their familiarity with this topic, their attitudes, norms, and perceptions of risks and benefits may change. In other words, these societal responses are highly dynamic, contextual, and varied in their intensity.

Respondents in a few mostly European studies believed that labeling and other forms of marketing and promotion could serve as one way of increasing awareness about forest biotechnology (Tsourgiannis et al., 2013, 2015, 2016; Kazana et al., 2015, 2016; Nonić et al., 2015). These studies showed that the majority of their respondents believed that labels should be required on any forest products involving the use of biotechnology, such as final products that originated from cisgenic or transgenic trees. Other European studies have discussed additional methods for increasing information about threats to forests (e.g., insect pests, pathogens, climate change) and increasing knowledge about potential biotechnological interventions for addressing these threats, including news coverage and social media attention (Tsourgiannis et al., 2013, 2015, 2016; Kazana et al., 2015, 2016; Nonić et al., 2015; Jepson and Arakelyan, 2017b). However, a couple of studies have shown that attitudes toward the use of biotechnology in trees and forests are extremely sensitive to informational messages and vulnerable to persuasion campaigns. For example, in an experiment involving samples of adults in the eastern United States and also students in Canadian and U.S. universities, Needham et al. (2016) found that acceptance of biotechnological interventions in trees and forests dropped dramatically (from 75–83 percent down to 40–44 percent) as soon as messages included any nega-

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

tive arguments (e.g., pejorative language) about this topic. Similarly, Hajjar et al. (2014) reported that acceptance changed for many of their Western Canadian respondents after being told that each potential intervention (including planting cisgenic or transgenic trees) would create either positive benefits or negative risks and other outcomes.

Given the lack of detailed knowledge about specific aspects of this topic and the potential malleability and instability of attitudes in response to informational or persuasive messages, trust in knowledgeable experts (e.g., forest agencies, scientists) is an important consideration for understanding perceptions and other responses (Brossard and Nisbet, 2006). Social trust is the willingness to rely on those responsible for making decisions or taking actions affecting public well-being (Connor and Siegrist, 2010). People may rely on trusted sources to assess complex or unknown issues. A number of studies in Europe and North America have shown that greater trust in forest managers (i.e., agencies) and scientists is associated with lower perceived risks, higher perceived benefits, more positive attitudes, and greater normative acceptance regarding the use of biotechnology in trees and forests (Neumann et al., 2007; Connor and Siegrist, 2010; Hajjar and Kozak, 2015; Needham et al., 2016; Jepson and Arakelyan, 2017a; Peterson St-Laurent et al., 2018). These trusted sources can use informational campaigns to increase knowledge that people can use for informing their support or opposition toward forest biotechnology in various contexts (e.g., private versus public land, plantations versus noncommercial forests; Strauss et al., 2017).

However, people may trust forest managers and scientists but may not always listen to them when the information provided conflicts with their own worldviews, beliefs, or values (Kahan et al., 2011; NRC, 2015). Values are abstract, enduring, and concerned with desirable end-states (e.g., safety, success) and modes of conduct (e.g., honesty, politeness). Values are basic modes of thinking that are shaped early in life by family or peers, are few in number and relatively stable, change slowly, and transcend situations and objects (Rokeach, 1973; Manfredo et al., 2004). There has been little research examining any potential direct relationships between these general values and more specific attitudes toward biotechnology in trees and forests. However, research has shown that a related, but different, concept called value orientations can be associated with these attitudes (Hajjar and Kozak, 2015; Needham et al., 2016; Peterson St-Laurent et al., 2018).

Value orientations reflect an expression of more general values and are revealed through the direction and strength of basic beliefs that an individual holds regarding more specific situations or issues (Manfredo et al., 2004). To measure value orientations toward forests, for example, Vaske and Donnelly (1999) asked individuals in the United States how strongly they agreed or disagreed with belief statements such as “the primary value of forests is to generate money and economic self-reliance for communities” and “forests have as much right to exist as people.” Patterns among these types of beliefs can be combined into value orientation continuums such as anthropocentric–biocentric, domination–mutualism (i.e., utilitarianism–affiliation or caring), and use–protection (Manfredo et al., 2004; Vaske and Manfredo, 2012). In the context of biotechnological interventions in trees and forests, Needham et al. (2016) reported that a representative sample of Americans with biocentric or environmental value orientations had slightly more positive attitudes toward using biotechnology to help trees resist chestnut blight and restore American chestnut forests than did those with anthropocentric or utilitarian value orientations. Hajjar and Kozak (2015) found that among a representative sample of Western Canadians, those with mixed or neutral environmental value orientations were slightly more accepting of biotechnology as a solution for addressing impacts of climate change on forests compared to those with more biocentric orientations. A more recent study in Western Canada showed that residents with anthropocentric value orientations were most supportive of using biotechnology in reforestation efforts in response to climate change (Peterson St-Laurent et al., 2018). Although these findings are mixed, they suggest that information campaigns, even from trusted sources, may have limited success in changing the attitudes of some people because these cognitions may be rooted in base values and value orientations that can be difficult to change.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

In addition, the discredited deficit model suggests that if people are given accurate information from sources that are considered to be objectively trustworthy and reliable (e.g., agencies, scientists), they would be more likely to overcome their limited knowledge and change their opinions to align with these sources (Davison et al., 1997). However, that is not how most people make decisions. In fact, more information about an issue does not always lead to greater knowledge or support, and it may even produce the opposite effect (Scheufele, 2006; NRC, 2015). For example, increased scientific information and communication from trusted sources may actually heighten risk perceptions, leading to more opposition toward the technology (Kellstedt et al., 2008). In addition, many people with limited knowledge about a complex issue (e.g., biotechnology) do not always base their decisions on new knowledge and information from trusted sources (i.e., cognitive reasoning). Instead, they often base their decisions on values, emotions, heuristics, schemas, and mental shortcuts, such as information from others who are like them and important to them (i.e., motivated reasoning; Brossard and Nisbet, 2006). In the context of biotechnology in forests, for example, critics may rely on intuitions and mental images (Blanke et al., 2015) such as “playing God,” “opposite of natural,” and “forest contamination,” whereas proponents may rely on notions of “technological progress,” “benevolent scientists,” or “wilderness as a managed garden.” This can lead to a confirmation bias, which is the tendency for people to seek information that reinforces their own values, supports what they already believe, and rejects disconfirming information even from objectively trustworthy sources (Kunda, 1990; Brossard and Nisbet, 2006; Scheufele, 2006; NRC, 2015).

Social and Ethical Values

As the research on public views outlined above suggests, developing biotechnology for use in trees and forests, especially in noncommercial and less intensively managed public forests, poses not only ecological and economic challenges, but also raises a range of social and ethical considerations. Some of these considerations directly relate to the provisioning of ecosystem services, including the perceived benefits to people and the environment (see Chapter 5 for a discussion of the complexity of ecosystem services), but some social and ethical considerations—especially those relating to intrinsic values of forests and social justice concerns—are not captured in ecosystem services. Although acknowledging that cultural components of ecosystem services provide a fairly broad and inclusive umbrella, this section explores social and ethical considerations as a complement to the ecosystem services framework. These considerations include intrinsic values, including the value of wildness, broad social influences, and social justice concerns.

Biotechnology and Forests’ Intrinsic Value

Because the idea of intrinsic value in nature is important for many conservationists (Justus et al., 2009), one consideration is what impact the use of biotechnology in forests might have on forests’ intrinsic value. The term “intrinsic value” can be used in different ways; the most relevant meaning here is intrinsic value understood as noninstrumental value, interpreted as the “value of things as ends in themselves, regardless of whether they are also useful as means to other ends” (Brennan and Lo, 2016). If a forest has intrinsic value in this sense, it has value in itself, above and beyond any use or service that it may provide human beings (as discussed in Chapter 2). Nonanthropocentric values are not easily captured in terms of traditional definitions and applications of ecosystem services (see, e.g., spiritual values in Box 4-1; see also Chapter 5), although the framework could recognize the benefit that some people experience in recognizing the intrinsic value of a species or ecosystem.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

The use of biotechnology in forests has the potential either to reduce or protect their intrinsic value. If biotechnology had the effect of making forests more easily available for human use and benefit, then it could undermine forests’ intrinsic value. However, whether it actually has this effect depends both on the purpose of the forests and the purpose of the particular biotechnology being used. Commercial forests are established primarily for consumptive use; they are likely to have significantly less intrinsic (noninstrumental) value than less intensively managed or noncommercial forests (see section “Biotechnology and Forests’ Naturalness or Wildness Value” below). Modifying the genome of a foundation or keystone tree species in a less intensively managed forest to increase resistance to an invasive pathogen or insect pest is not a way of using the forest for human benefit. Whitebark pine, for instance, though foundational in its ecosystem, has little commercial use; biotechnological changes to increase resistance to invasive blister rust would not make them more easily available for human benefit.

Biotechnology and Forests’ Naturalness or Wildness Value

The use of biotechnology, though, may still have implications for intrinsic value, depending on what is actually being intrinsically valued about the forest. One important way in which forests may be valued intrinsically is in terms of their wildness or naturalness. It should be noted that the meaning, existence, and value of “wild,” “wilderness,” and “natural” in environments such as forests have been widely contested. A number of scholars have argued that the idea of a valuable “wildness” when located in environments perceived as wilderness or otherwise “set apart” from people is historically and culturally specific, based on a problematic dualism between humans and nature, and can lead to devastating impacts on indigenous peoples who occupy such “wild” places (e.g., Denevan, 1992; Cronon, 1995; Callicott and Nelson, 1998; DeLuca and Demo, 2001; Nelson and Callicott, 2008). However, the value of “naturalness” has already played a significant role in debates about biotechnology, and it can be expected that “wildness” will be important in thinking about

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

the use of biotechnology in less managed or unmanaged forests. Where wildness or naturalness are intrinsically valued, there may be significant concerns that biotechnology could reduce this intrinsic value. One widely expressed concern about the use of biotechnology, as some of the research on public opinions discussed above suggests, is that it is considered to be “tampering with nature” or “unnatural” (Sjöberg, 2004; Hajjar and Kozak, 2015; Jepson and Arakelyan, 2017b; Lull and Scheufele, 2017). If biotechnology is seen as extending new, or more intense, human “tampering” into forests previously valued for their naturalness, then biotechnology could be seen as undermining a forests’ value in this sense. On the other hand, if intrinsic value is (in part at least) based on the continued natural or wild existence of a particular threatened tree species or population, biological diversity, or the continued health of the entire forest ecosystem, then the use of biotechnology for forest health may be regarded as protecting intrinsic value.

When considering biotechnology use in less intensively managed forests (e.g., public, noncommercial), there may be different kinds of concerns about unnaturalness. One is a broader concern about the “unnaturalness” of biotechnological processes, the kind of concern that has also been expressed about the use of biotechnology in agricultural crops. Here, “unnaturalness” denotes “whether it could have taken place without human beings” (Siipi, 2015:810). This understanding of “naturalness” may partly explain the findings of Jepson and Arakelyan (2017a), noted above, that in the case of ash dieback in the United Kingdom, cisgenic approaches were preferred over transgenic methods by the residents surveyed. Cisgenesis might be regarded as more “natural” in the sense that it is more likely to occur without human intervention than genetic modification through transgenesis.

However, “unnaturalness” in forests may also refer to wildness understood somewhat differently. Hettinger and Throop (1999:12) defined wildness in a place or thing as “something is wild in a certain respect to the extent that it is not humanized in that respect. An entity is humanized in the degree to which it is influenced, altered or controlled by humans.” In the case of forests, wildness might refer to many characteristics such as wild origins (humans have not chosen which trees are planted where, but a process of “natural” seed distribution has created the forest); wild composition (humans have not decided which species are found where); and wild processes (spontaneous evolutionary and ecological processes are continuing without human intervention; humans are not thinning or felling trees, removing dead wood or making other management decisions that control forest processes). The use of biotechnology may be thought to undermine wildness in forests in any or all of these senses by disrupting the perception of wild origins, composition, and processes. Given that biotechnology has the effect of extending human management, influence, and intention, those forests would lose some of their perceived wildness by becoming more entwined with human action.

Any decision to use biotechnology in forests is guided by human preference for a particular future for a forest (e.g., forests should continue to contain specific species), even if that preference is directed at protecting or promoting forest health. To use Hettinger and Throop’s (1999) terms, biotechnology is intended to influence and alter the forest and could be interpreted as a form of human control of a forest ecosystem that previously, in some sense, was “self-directing” or “autonomous.” The use of biotechnology may also affect wildness in the more specific senses mentioned above. For instance, transgenic or genome-edited trees of species chosen by humans are likely to be planted in places selected by humans and for some period at least managed and monitored by humans, which could be understood to reduce wildness in terms of origin, composition, and process value. The use of biotechnology is also a human intervention in the “natural” evolutionary trajectory of the forest. Although the use of biotechnology may promote forest health, it may nonetheless be perceived as diminishing the wildness value of forests. In this sense, debate about forest biotechnology is likely to resemble that of ecological restoration, where concerns have been expressed that the human origin of an ecological restoration makes it less valuable than the original ecosystem, even if the restored system is flourishing and healthy (Elliot, 1982; Katz, 1992).

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

On the other hand, threats to forests that biotechnology may counter are predominantly of human origin (e.g., invasive insect pests and pathogens transported by people and native insect pests and pathogens extending their range because of human influences on climate). Given that these changes are also signs of human influence, forest wildness may already be seen as reduced, if not undermined significantly. Doing nothing to counter such anthropogenic threats may result in the loss of particular populations or entire species, with significant effects on forest ecosystems that at least in some senses (e.g., species composition) also mean a loss of wildness. In addition, other practices that might protect forest health, such as selective breeding, seem to pose rather similar threats to wildness as biotechnology because they also involve the selection of particular genotypes, the decision to plant trees in particular places, continued monitoring of the trees, and so on. So, although the use of biotechnology in forests may diminish their perceived wildness value, alternative options (including, perhaps, no action at all) also reduce wildness, albeit to varying degrees and in different ways.

This can be seen particularly clearly in the case of the whitebark pine’s status as “symbols of the primeval forest, the wilderness, and the forces of nature” (Tomback and Achuff, 2010:201). If humans were to intervene in the genome of the whitebark pine, wildness would be reduced in one sense; the genome of all future members of the species would now be influenced by decisions made by humans, which seems to make them less “primeval” and less symbolic of the “forces of nature.” On the other hand, whitebark pines are already being threatened by invasive blister rust, introduced by humans planting American white pines that had been grown in Europe and then brought into the United States; without intervention, the whitebark pine may be extirpated in many places, or driven to extinction. The difficulty lies in deciding how to evaluate whether forests devoid or greatly reduced of whitebark pines due to human-driven invasive species and climate change would have more wildness value than forests populated by biotech whitebark pines.

Biotechnology and the Intrinsic Value of Forest Species, Ecosystems, and Biodiversity

Many environmental philosophers, conservationists, and conservation biologists claim that species (e.g., Soulé, 1985; Rolston, 1988; Smith, 2016), ecosystems (e.g., Leopold, 1949; Callicott, 1989), and biological diversity (Soulé, 1985; UN, 1992) have intrinsic value. These claims may have different justifications. For instance, it may be argued that species’ intrinsic value rests on their long evolutionary history and potential (Soulé, 1985), or alternatively rests on the grounds that species have interests and a good of their own that are of moral significance and should be respected (Johnson, 1991). Although such claims are contested and controversial (Sandler, 2012), they are likely to feature in future debates about the use of biotechnology in forests, alongside discussions of ecosystem services.

These value positions suggest that at least some uses of biotechnology to protect or promote forest health are likely to be viewed by some constituents as positively impacting or enhancing intrinsic value. The use of biotechnology to restore the American chestnut, for example, could be understood as protecting both the intrinsic value of this species and forest ecosystems by improving their health, and also as promoting intrinsically valuable forest biodiversity by reintroducing a species on which a wide variety of other organisms depend (Powell, 2016). Similar arguments might be made for the other tree species on which the committee has focused in this report. Positive interpretations of this kind are supported by Needham and colleagues’ (2016) finding that having a stronger biocentric or environmental value orientation tended to underpin a more positive attitude toward the use of biotechnology to help restore American chestnut forests.

However, this should not be taken to imply general acceptance of the use of biotechnology to promote forest health. In some cases, an application of biotechnology could present risks to certain intrinsic values, even as it protects other values. For instance, the use of biotechnology in

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

a tree species to protect it against an invasive insect might threaten the survival of other native or endemic insect species. In addition, those who defend the intrinsic value of species, ecosystems, or biodiversity may also accept the intrinsic value of naturalness or wildness (Leopold, 1949; Soulé, 1985), meaning that the use of biotechnology for forest health could entail choosing between environmental values such as species preservation and wildness protection. Given that the use of biotechnology in forests may undermine some values while enhancing others, each possible use of biotechnology in forests is likely to need its own individual ethical case analysis (Sandler, 2018).

This potential choice brings into focus a set of broader social and ethical debates about the use of new technologies in environmental conservation. The pervasiveness of ecological impacts from forces such as climate change and invasive species means that traditional conservation strategies, including setting aside nature reserves and restoring species to habitats within their historic ranges, become significantly less likely to achieve goals such as species protection (Minteer and Collins, 2012; Sandler, 2013, 2018). To protect species and reduce biodiversity loss may then require new interventionist and managerial conservation strategies, such as the use of biotechnology. In addition to potentially negatively impacting perceptions of wildness, the use of such new technologies changes the nature of traditional conservation practices, thus raising a variety of challenges about the broader social influences of technology.

Challenges Raised by Broader Social Influences of Biotechnology

Biotechnologies developed to protect and promote forest health target particular genes in specific species for particular purposes. Although these targets may be specific and narrow, many social scientists have argued that the uses of new technologies almost always have much wider social and cultural impacts than their immediate target (Johnson and Wetmore, 2009; Slovic, 2010). Winner (2010:6) maintains that “technologies are not merely aids to human activity, but also powerful forces acting to reshape that activity and its meaning.” Technologies such as the automobile or the cellular telephone transformed societies as they were adopted, changing people’s sense of identity, the nature of human relationships and interactions, the nature of and access to employment, and people’s everyday habits. Likewise, the use of biotechnology for conservation purposes could have much broader societal effects, including the potential for reshaping some conservation purposes and practices, effects less likely to follow from the use of more traditional techniques such as tree breeding. For instance, the use of biotechnology for conservation purposes could promote a shift in the focus of conservation from more traditional calls to change human behaviors in the environment, or attempts to separate places and species from undue human impacts (e.g., by creating nature reserves), to much more managerial and interventionist strategies involving altering species and ecosystems to better fit into a human adapted world (Gamborg and Sandøe, 2010; Sandler, 2018). One particular concern here is that the use of biotechnology for forest health could have the effect of making the adoption of biotechnology seem more routine, thus serving as a perceived portal or “Trojan horse” for future biotech modifications in forests or other environments for very different—and less altruistic—purposes (Smolker, 2018).

Scholars in the social studies of science and technology have focused on processes and potential institutions to understand such potential and complex impacts through innovations in:

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Such processes may include attention to risks and benefits—the primary focus of U.S. regulatory oversight of biotech plants—but also expand to consider a much broader and diverse set of values in the context of uncertainty.

Social Justice Considerations in the Use of Biotechnology for Forest Health

The use of biotechnology for forest health also potentially raises social justice challenges, which may be overshadowed by analyses focusing on the services that forest ecosystems provide, rather than how the benefits, costs, and risks derived from those services are distributed. These social justice challenges relate to (1) distributive justice, defined as “the political processes and structures that affect the distribution of benefits and burdens in societies” (Lamont and Favor, 2017); and (2) procedural justice, defined as “the justice of the procedures that might be used to determine how benefits and burdens of various kinds are allocated to people” (Miller, 2017). Given that the use of biotechnology in forests affects the future of forests—and therefore of humans—over the long term, social justice challenges also extend beyond present generations to include future generations, raising intergenerational justice challenges as well.

Distributive Justice. The use of biotechnology in forests raises possible issues of distributive justice. The most obvious justice concerns are likely to be raised where some individuals or groups bear a disproportionate share of the risks or harms from the use of biotechnology in forests, but receive few or no benefits. “Risks” and “harms” here do not primarily refer to risks to human health; relevant risks, for example, could be to the livelihood or cultural practices of forest-dependent communities. If the use of biotechnology in noncommercial forests reduced tourism, there might be a negative impact on those who depend on tourism for their livelihood (though possibly no worse than if the forest were severely affected by an insect or pathogen). Alternatively, stakeholders seeking to restore a tree species such as the American chestnut might benefit from the introduction of blight-resistant transgenic American chestnut trees, whereas stakeholders who view any genetic modification of a forest species as reducing its wildness will bear the harm. In this particular example, perceptions and values drive the distribution of harm and benefit more so than geography, race, or class.

Indigenous populations who have spiritual relationships with, and value for, particular forests and tree species are likely to be significantly affected by the use of biotechnology in noncommercial forests (Nilausen et al., 2016; see also Box 4-1). For example, black ash (Fraxinus nigra) has special significance for indigenous peoples in the Great Lakes region, especially for basket making (Poland et al., 2017). Although black ash is seriously threatened by the emerald ash borer, the use of biotechnology to increase resistance in black ash trees might significantly change the relationship indigenous peoples have to this species. Relatedly, recent research on the potential restoration of the American chestnut tree draws on interviews with Haudenosaunee community members and participant observation of tribal meetings. Barnhill-Dilling (2018) acknowledged great diversity in perspectives among the indigenous people with whom she interacted, but reports several themes relevant to this discussion; the committee heard similar information in one of its webinars (Dockry, 2018; McManama, 2018; Patterson, 2018). First, traditions of nonintervention in natural processes (Nelson, 2008) question the wisdom of attempting to counteract the effects of the chestnut blight altogether. Second, cultural and medicinal practices that used to involve the American chestnut tree are viewed as unlikely to be restored with a transgenic or hybrid tree. Third, disrespect and abuse of native peoples by Western scientists (Sikes, 2006; Smith, 2013) has created a culture of mistrust that fosters skepticism of scientific innovations even when they are presented as beneficial. Fourth, in a period of increased attention to indigenous cultural restoration, a narrow focus on the restoration of a single tree species is experienced by some tribal members as marginal, if not irrelevant (also see Higgs, 2005; Kimmerer, 2011). Fifth, and most broadly, in some indigenous communities,

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

genetic engineering has come to be viewed as violating tribal sovereignty, self-determination, and the natural order (also see Harry, 2001; Roberts, 2005; Antoine, 2014; Francis, 2015; IEN, 2016) and, as noted in Box 4-1, might be interpreted as violating indigenous peoples’ rights. However, it is important not to overgeneralize. Barnhill-Dilling (2018) also reported that some indigenous representatives see chestnut blight as a destructive force and welcome the potential for a transgenic tree to restore both ecological integrity and cultural practices related to woodworking and eating chestnuts. Thus, the distribution of risks and benefits across cultural, social, and sovereign boundaries introduces great complexity in considering the social justice dimensions of forest biotechnology.

Distributive justice presents the challenge of considering not only potential risks, harms, and benefits from the use of biotechnology in forests, but also the ways in which those risks, harms, and benefits are distributed across populations and individuals. However, it is also important to note that the existing threat—such as from new insect pests or pathogens—to which any proposed biotechnology is responding also generates risks, harms, and benefits (e.g., employment from thinning diseased forests or a new desired species composition after the pest has gone through the forest) distributed unevenly across populations and individuals. There is not a no-risk scenario for the cases in which biotechnology use is being considered.

Procedural Justice. The challenge of procedural justice is to ensure that those who are likely to be affected meaningfully participate in decision making about the use of biotechnology in forests; this requires inclusiveness in consultation and decision-making procedures. Including all those who are likely to be affected generates particular challenges in the case of a modification that is designed to spread in the environment and across social and political boundaries. Biotechnology used in one forest is eventually likely to reach forests (and those living in, dependent on, or visiting those forests) at a considerable distance from where trees were initially planted. Therefore, consulting only those people local to a particular proposed biotechnological use appears to be too limited.

Procedural justice also requires recognition of the standing of particular cultural groups who will be affected. For example, indigenous groups should be meaningfully and fairly included in consultations about uses of biotechnology that may affect their forests and in ways that allow for the sharing of indigenous knowledge (McGregor, 2002) and that recognize tribal sovereignty, cultural practices, and values, even where those values diverge from the values of other communities and individuals who may also be affected (Barnhill-Dilling, 2018). Consultation is already required if federal policies affect indigenous communities (see the example in Box 6-1).8

Attending to procedural justice is not a recipe for avoiding controversy. In fact, expanding the number of individuals and groups meaningfully consulted is unlikely to make consensus any easier to achieve. There will likely be objections to any decisions ultimately taken. However, ensuring procedural justice allows those with authority to explain how and why particular values were prioritized, how the steps toward decision making in each case were made, and who was responsible for them, therefore displaying transparency in the decision-making procedure. Put simply, procedural justice helps ensure fairness even if outcomes are unlikely to satisfy all members of various publics.

Intergenerational Justice. Concerns about social justice extend beyond those currently alive to include future generations of human beings. Many tree species are long-lived, with life spans exceeding many human generations. Whitebark pines, for instance, do not reach reproductive age until they are at least 20 years old, may not reach maximum cone production until they are 100, and can live for more than 1,000 years. The use of biotechnology in trees that are likely to outlive those who planted them, and that will affect the species composition of forests for centuries, clearly

___________________

8 Examples of consultation models and practices with indigenous communities include Whyte et al. (2014), Farley et al. (2015), and Norton-Smith (2016).

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

has implications for future generations. In terms of distributive justice, if present generations could be expected to benefit from the use of biotechnology in forests, whereas future generations bore a disproportionate share of the risks and costs, this would present an issue of intergenerational injustice. However, the benefits, risks, and costs may not split this way at all; it is plausible that present generations would bear the costs of developing, breeding, and planting trees generated using biotechnology, whereas future generations would benefit from the resulting mature trees and healthy forests.

This issue is particularly challenging given the uncertainties of the effects of biotechnology over long timescales, limited knowledge about the future trajectories of current and new forest pests and climate change, and the fact that future generations cannot directly be consulted about their values and preferences with respect to the use of biotechnology. In the language of procedural justice, it is difficult to imagine a procedure that integrates the perspectives and concerns of publics of the future, although a number of ways of integrating such concerns into democratic systems have been proposed (e.g. Thompson, 2010; González-Ricoy and Gosseries, 2016).

Many other human impacts on the environment will affect things over the long term, not just the use of biotechnology in forests. Tree species have approached extinction in the past due to new pests and have recovered (Booth et al., 2012). It is possible that, in some respects, the use of biotechnology for forest health would make less of a long-term impact than the extirpation of populations or even species extinction that could have been averted by the use of such technology. However, the long-term nature of this form of biotechnology is highlighted by the transformation of such long-lived organisms as trees (in comparison with the planting of annual crops, for instance). Frameworks that focus on “sustainability” or the preservation of options for future generations may offer instructive insight (e.g., Hauser et al., 2014).

CONCLUSIONS AND RECOMMENDATIONS

Because trees are long-lived species that often exist in minimally managed or unmanaged environments, there are a number of ecological, economic, social, and ethical considerations that pertain to the use of biotechnology in forest trees that are not as applicable to other biotech products, such as agricultural crops. To be an effective tool in the mitigation of forest health threats, these various considerations need to be taken into account when making decisions about the deployment of a tree with biotech resistance to insect pests or pathogens.

Conclusion: Trees with resistance introduced via biotechnology will have to survive until maturity and reproduce in order to pass resistant traits on to the next generation.

Because forest trees are in minimally managed or unmanaged environments, biotech trees with resistance to pests will have to be genetically fit in their respective environments and capable of competing with other plant species to become established. They will also have to be able to convert the resistance trait into future generations without expressing additional traits, such as high fecundity and rapid growth rate, which could lead to invasiveness.

Recommendation: Research should address whether resistance imparted to tree species through a genetic change will be sufficient to persist in trees that are expected to live for decades to centuries as progenitors of future generations.

Conclusion: The importance of managing and conserving standing genetic variation to sustain the health of forests cannot be overstated.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

The postglacial expansion of tree species out of the glacial refugia has shaped genetic variation in forest populations and enabled local adaptation that appears to be pervasive in widely distributed species. In this context, it is worth considering the adaptability and vulnerability of populations under future climates. Fitzpatrick and Keller (2015) demonstrated that the vulnerability could be measured in terms of genetic offset, a metric that identifies populations within the species’ distribution where local adaptation gene × environment relationship, is most likely to be disrupted. For example, their modeling has shown that for the widely distributed boreal tree, Populus balsamifera, the genetic offset is the greatest along the northern range edge (Fitzpatrick and Keller, 2015). Identifying spatial regions most vulnerable to genetic offset under future environmental conditions can therefore lead to better conservation and management practices.

Recommendation: The deployment of any biotechnological solution with the goal of preserving forest health should be preceded by developing a reasonable understanding in the target species of (a) rangewide patterns of distribution of standing genetic variation including in the putative glacial refugia, if known; (b) magnitude of local adaptation (gene × environment relationships); and (c) identification of spatial regions that are vulnerable to genetic offset.

Conclusion: The public sector will be best positioned to lead development of biotech trees because of the public-good aspect of forest health and the intention for the spread of a biotech tree through a forest ecosystem.

The role of the public sector (including government and nonprofit entities such as private foundations) arises out of the likelihood that the private sector will not invest in the protection of forest health because it cannot fully capture the benefits that may accrue and because it will not be able to restrict access to a tree that is released with the intention that it propagate freely. Without the expectation of market revenue sufficient to justify the costs of development, the private sector will not have sufficient incentive to invest its resources. Beyond this market failure, the justification for use of public funds arises from the nonmarket benefits of healthy forests.

Conclusion: The relatively long time required for the development of a biotech tree may adversely affect the incentive for both private- and public-sector investment.

The costs of development of a biotech tree (or, indeed, any tree bred or designed for pest resistance) will be incurred up front and the benefits will follow years later. Such a difference in the timing makes investment with a long time horizon problematic. Compared to the private sector, the public sector can have greater patience when significant public benefits are forthcoming.

Conclusion: Few studies of public attitudes toward biotechnology to address forest health threats have yet been carried out in the United States. However, there has been a small handful of studies on the topic, especially in Canada and Europe. The limited data indicate that while some individuals and groups are very concerned about possible deployment of biotechnology in forests, attitudes toward the uses of biotechnology examined in these studies are somewhat positive, especially where threats to forests are severe.

Compared to the number of studies examining societal views about the use of biotechnology in agriculture and food, there have been few studies about how people think and feel about the use of biotechnology in trees to address forest health threats. Most studies have reported that the majority of study participants supported some biotechnological approaches, which were often viewed as more acceptable than doing nothing to address severe threats to forest health.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Conclusion: Existing research indicates that public knowledge and understanding about the use of biotechnology in forests is low, suggesting that current attitudes may be unstable and liable to change with more information. The power of such information to influence attitudes is mediated by the perceived trust of the sources of information, deliberation about the topic, and the alignment of new information with deep value orientations.

The lack of detailed knowledge by most members of the public about biotechnology, forest health, and the biotech and nonbiotech tools that could be used to address forest health means that attitudes toward the use of biotechnology in forest trees are extremely sensitive to informational messages and vulnerable to persuasion campaigns. Information delivered by trusted knowledgeable experts (e.g., forest agencies, scientists) may influence attitudes, but information campaigns, even from trusted sources, may have limited success in changing attitudes depending on people’s values and value orientations.

Conclusion: Some important ethical questions raised by deploying biotechnology in noncommercial forests fall outside any evaluation of changes in ecosystem services.

The use of biotechnology may negatively affect perceptions of noncommercial forests’ wildness or naturalness. Conversely, the use of biotechnology may protect forests, in terms of preventing the loss of valuable species, ecosystems, and biodiversity. It may also affect the spiritual interactions some individuals and cultural groups have with forests. In some cases, the use of biotechnology may mean that protecting one value, such as a threatened species, means sacrificing another value, such as wildness. These potential trade-offs indicate the need for case-specific ethical management assessments that take into account the different values at stake both in any proposed use of biotechnology, or in not intervening at all.

Recommendation: More studies of societal responses to the use of biotechnology to address forest health threats in the United States are needed. Such studies might investigate (1) the responses of different social and cultural groups to the deployment of biotechnology in forests, (2) the stability and consistency of attitudes toward different applications of biotechnology in a range of circumstances, (3) differences in attitudes toward biotechnology strategies (e.g., cisgenesis, transgenesis, genome editing), (4) the relationship between deeper value orientations and attitudes toward biotechnology, and (5) how people consider trade-offs between values such as wildness and species protection.

Conclusion: The use of biotechnology for forest health, especially in noncommercial forests, raises broad questions about the social impacts of technological change on society, in particular, how conservation is understood and practiced, and how far biotechnological interventions presage a change to more interventionist management of forests.

The automobile and the cellular telephone transformed societies, and the use of biotechnology for conservation purposes could also have broad societal effects. It has the potential to reshape conservation purposes and practices; for example, it could promote a shift from calls to change human behaviors in order to save the environment, to more managerial and interventionist strategies involving altering species and ecosystems to better fit into a human-adapted world. Understanding the complex impacts on society of using biotechnology in trees in minimally managed and unmanaged environments will require more study in areas of governance and public engagement.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Conclusion: The use of biotechnology for forest health raises social justice questions, both in terms of the distribution of risks, harms, and benefits across individuals and groups through time, and in terms of the procedures used to make decisions about whether, when, and where to deploy the technology. Indigenous communities may be particularly affected by these decisions. Given the longevity of trees, the use of biotechnology for forest health (or the decision not to use it) will have significant impacts on future generations.

Distributive justice is concerned with potential risks, harms, and benefits and the ways in which those risks, harms, and benefits are distributed across populations and individuals. These concerns apply to the use of biotechnology in forests as well as to the threats posed by insect pests and pathogens to forest health. Procedural justice seeks to ensure that those who are likely to be affected meaningfully participate in decision making about the use of biotechnology in forests; this requires inclusiveness in consultation and decision-making procedures. However, it does not guarantee that all parties will be satisfied with the outcome. Intergenerational justice recognizes that concerns about social justice extend to subsequent generations, which is particularly pertinent to the use of biotechnology in trees since many trees will outlive those who create and plant parent trees with resistance to forest pests. It is difficult to predict how risks, benefits, and costs will be distributed among generations because of the uncertainties of the effects of biotechnology over long timescales, limited knowledge about the future trajectories of current and new forest pests and climate change, and the fact that future generations cannot be directly consulted about their values and preferences with respect to the use of biotechnology. This uncertainty is similar for many human impacts on the environment.

Recommendation: Respectful, deliberative, transparent, and inclusive processes of engaging with people should be developed and deployed, both to increase understanding of forest health threats and to uncover complex public responses to any potential interventions, including those involving biotechnology. These processes, which may include surveys, focus groups, town hall meetings, science cafés, and other methods, should contribute to decision making that respects diverse sources of knowledge, values, and perspectives.

REFERENCES

Aitken, S.N., and J.B. Bemmels. 2016. Time to get moving: Assisted gene flow of forest trees. Evolutionary Applications 9(1):271–290.

Aitken, S.N., and M.C. Whitlock. 2013. Assisted gene flow to facilitate local adaptation to climate change. Annual Review of Ecology, Evolution, and Systematics 44:367–388.

Aitken, S.N., S. Yeaman, J.A. Holiday, T. Wang, S. Curtis-McLane. 2008. Adaptation, migration or extirpation: Climate change outcomes for tree populations. Evolutionary Applications 1(1):95–111.

Alberto, F.J., S.N. Aitken, R. Alía, S.C. González-Martínez, H. Hänninen, A. Kremer, F. Lefèvre, T. Lenormand, S. Yeaman, R. Whetten, and O. Savolainen. 2013. Potential for evolutionary responses to climate change—evidence from tree populations. Global Change Biology 19:1645–1661.

Antoine, D. 2014. Standing Against GMOs Is Standing for Sovereignty. Indian Country Today Media Network, March 5. Available at https://newsmaven.io/indiancountrytoday/archive/standing-against-gmos-is-standing-for-sovereignty-pICTeCjyT0yiJloCO_KT7w. Accessed September 18, 2018.

Arnold, M.L. 1992. Natural hybridization as an evolutionary process. Annual Review of Ecology and Systematics 23:237–261.

Barnhill-Dilling, S.K. 2018. Engineered Governance: Public Engagement as a Path to Inclusive Environmental Governance in Genetically Engineered Trees. Ph.D. dissertation, North Carolina State University. Available at https://repository.lib.ncsu.edu/handle/1840.20/35317. Accessed November 21, 2018.

Barnosky, C.W., P.M. Anderson, and P.J. Bartlein. 1987. The northwestern US during deglaciation; vegetational history and paleoclimatic implications. Pp. 289–321 in North America and Adjacent Oceans During the Last Deglaciation, Vol. 3, W.F. Ruddiman and H.E. Wright, Jr., eds. Boulder, CO: Geological Society of America.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Barrett, R.D., and D. Schluter. 2008. Adaptation from standing genetic variation. Trends in Ecology & Evolution 23(1):38–44.

BEA (U.S. Bureau of Economic Analysis). 2018. Industry Facts. Available at https://www.bea.gov. Accessed May 22, 2018.

Blanke, S., F. Van Breusegem, G. De Jaeger, J. Braeckman, and M. Van Montagu. 2015. Fatal attraction: The intuitive appeal of GMO opposition. Trends in Plant Science 20(7):414-418.

BLS (U.S. Bureau of Labor Statistics). 2018. Employment by Sector. Available at https://www.bls.gov. Accessed May 22, 2018.

Bonan, G.B. 1999. Frost followed the plow: Impacts of deforestation on the climate of the United States. Ecological Applications 9(4):1305–1315.

Booth, A., and H.M. Jacobs. 1990. Ties that bind: Native American beliefs as a foundation for environmental consciousness. Environmental Ethics 12(1):27–43.

Booth, R.K., S. Brewer, M. Blaauw, T.A. Minckley, and S.T. Jackson. 2012. Decomposing the mid-Holocene Tsuga decline in eastern North America. Ecology 93(8):1841–1852.

Bower, A.D., and S.N. Aitken. 2008. Ecological genetics and seed transfer guidelines for Pinus albicaulis (Pinaceae). American Journal of Botany 95(1):66–76.

Boyd, J., P. Ringold, A. Krupnick, R.J. Johnston, M.A. Weber, and K. Hall. 2016. Ecosystem services indicators: Improving the linkage between biophysical and economic analysis. International Review of Environmental and Resource Economics 8(3–4):359–443.

Brennan, A., and Y.S. Lo. 2016. Environmental ethics. In The Stanford Encyclopedia of Philosophy (winter edition), E.N. Zalta, ed. Available at https://plato.stanford.edu/archives/win2016/entries/ethics-environmental. Accessed April 15, 2018.

Brockerhoff, E.G., L. Barbaro, B. Castagneyrol, D.I. Forrester, B. Gardiner, J.R. González-Olabarria, P.O.B. Lyver, N. Meurisse, A. Oxbrough, H. Taki, I.D. Thompson, F. van der Plas, and H. Jactel. 2017. Forest biodiversity, ecosystem functioning and the provision of ecosystem services. Biodiversity and Conservation 26:3005–3035.

Brossard, D., and M.C. Nisbet. 2006. Deference to scientific authority among a low information public: Understanding U.S. opinion on agricultural biotechnology. International Journal of Public Opinion Research 19(1):24–52.

Brown, J.E. 1985. The Spiritual Legacy of the American Indian. New York: Crossroad.

Callicott, J.B., 1989. In Defense of the Land Ethic: Essays in Environmental Philosophy. Albany, NY: State University of New York Press.

Callicott, J.B., and M.P. Nelson, eds. 1998. The Great New Wilderness Debate. Athens, GA: University of Georgia Press.

Cartwright, C. 2018. Whitebark pine (Pinus albicaulis) field screening for blister rust resistance in British Columbia: Germination results. Pp. 81–87 in Proceedings of the IUFRO Joint Conference: Genetics of Five-Needle Pines, Rusts of Forest Trees, and Strobusphere, A.W. Schoettle, R.A. Sniezko, and J.T. Kliejunas, eds. Fort Collins, CO: U.S. Forest Service.

Cartwright, C., N. Ukrainetz, and M. Murray. 2016. Whitebark Pine Field Screening for Blister Rust Resistance: Establishment Report–2015/2016. Province of British Columbia.

Chang, W., V.A. Lantz, and D.A. MacLean. 2009. Public attitudes about forest pest outbreaks and control: Case studies in two Canadian provinces. Forest Ecology and Management 257(4):1333–1343.

Clark, J.S. 2018. Our charge: Forest health measures for evaluating risk from release of biotech trees. Presentation to the National Academies of Sciences, Engineering, and Medicine Committee on the Potential for Biotechnology to Address Forest Health, February 8, Washington, DC.

Clark, W. 2011. Clarifying the spiritual value of forests and their role in sustainable forest management. Journal for the Study of Religion, Nature and Culture 5(1):18–38.

Cohen, J. 2018. Federal appeals court hears CRISPR patent dispute. Science, April 30. Available at http://www.sciencemag.org/news/2018/04/federal-appeals-court-hears-crispr-patent-dispute. Accessed September 18, 2018.

Connor, M., and M. Siegrist. 2010. Factors influencing people’s acceptance of gene technology: The role of knowledge, health expectations, naturalness, and social trust. Science Communication 32(4):514–538.

Cooper, N., E. Brady, H. Steen, and R. Bryce. 2016. Aesthetic and spiritual values of ecosystems: Recognising the ontological and axiological plurality of cultural ecosystem “services.” Ecosystem Services 21(Part B):218–229.

Costa-Font, M., J.M. Gil, and W.B. Traill. 2008. Consumer acceptance, valuation of and attitudes towards genetically modified food: Review and implications for food policy. Food Policy 33(2):99–111.

Costanza, R., R. de Groot, P. Sutton, S. van der Ploeg, S.J. Anderson, I. Kubiszewski, S. Farber, R.K. Turner. 2014. Changes in the global value of ecosystem services. Global Environmental Change 26:152–158.

Cronon, W. 1995. The trouble with wilderness or getting back to the wrong nature. Pp. 69–90 Uncommon Ground: Toward Reinventing Nature, W. Cronon, ed. New York: W.W. Norton.

Crowley, K.F., G.M. Lovett, M.A. Arthur, and K.C. Weathers. 2016. Long-term effects of pest-induced tree species change on carbon and nitrogen cycling in northeastern US forests: a modeling analysis. Forest Ecology and Management 372:269–290.

Davison, A., I. Burns, and R. Schibeci. 1997. Problematic publics: A critical review of surveys of public attitudes to biotechnology. Science, Technology, and Human Values 22(3):317–348.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Delborne, J.A., J. Schneider, R. Bal, S. Cozzens, and R. Worthington. 2013. Policy pathways, policy networks, and citizen deliberation: Disseminating the results of World Wide Views on Global Warming in the USA. Science and Public Policy 40(3):378–392.

Delcourt, H.R., and P.A. Delcourt. 1984. Ice age haven for hardwoods. Natural History 84:22–28.

DeLuca, K., and A. Demo. 2001. Imagining nature and erasing class and race: Carleton Watkins, John Muir, and the construction of wilderness. Environmental History 6(4):541–560.

Denevan, W.M. 1992. The pristine myth: The landscape of the Americas. Annals of the Association of American Geographers. 82(3):369–385.

Dockry, M.J. 2018. Indigenous perspectives on the use of biotechnology in forests. Webinar presentation to the National Academies of Sciences, Engineering, and Medicine’s Committee on the Potential for Biotechnology to Address Forest Health, March 2.

Dumroese, R.K., M.I. Williams, J.A. Stanturf, and J.B.S. Clair. 2015. Considerations for restoring temperate forests of tomorrow: Forest restoration, assisted migration, and bioengineering. New Forests 46(5–6):947–964.

Eagly, A.H., and S. Chaiken. 1993. The Psychology of Attitudes. Belmont, CA: Wadsworth.

Eckert, C.G., K.E. Samis, and S.C. Lougheed. 2008. Genetic variation across species’ geographical ranges: The central–marginal hypothesis and beyond. Molecular Ecology 17(5):1170–1188.

Elliot, R. 1982. Faking nature. Inquiry 25(1):81–93.

Ellstrand, N.C. 1992. Gene flow by pollen: Implications for plant conservation genetics. Oikos 63(1):77–86.

EPA (U.S. Environmental Protection Agency). 2014. Guidelines for Preparing Economic Analyses. Washington, DC: EPA.

FAO (United Nations Food and Agriculture Organization). 1998. Economic and Environment Accounting for Forestry: Status and Current Efforts. Rome, Italy: FAO.

FAO. 2016. Global Forest Resources Assessment 2015: How Are the World’s Forests Changing? Second edition. Rome, Italy: FAO.

Farley, C., T. Ellersick, and C. Jasper, eds. 2015. Forest Service Research and Development Tribal Engagement Roadmap. Available at https://www.fs.fed.us/research/docs/tribal-engagement/consultation/roadmap.pdf. Accessed November 17, 2018.

Feurtey, A., A. Cornille, J.A. Shykoff, A. Snirc, and T. Giraud. 2017. Crop to wild gene flow and its fitness consequences for a wild fruit tree: Towards a comprehensive conservation strategy of the wild apple in Europe. Evolutionary Applications 10(2):180–188.

Finucane, M.L., and J.L. Holup. 2005. Psychosocial and cultural factors affecting the perceived risk of genetically modified food: An overview of the literature. Social Science and Medicine 60(7):1603–1612.

Fishbein, M., and I. Ajzen. 2010. Predicting and Changing Behavior: The Reasoned Action Approach. New York: Psychology Press.

Fitzpatrick, M.C., and S.R. Keller. 2015. Ecological genomics meets community-level modelling of biodiversity: Mapping the genomic landscape of current and future environmental adaptation. Ecology Letters 18(1):1–16.

Flint, C.G. 2006. Community perspectives on spruce beetle impacts on the Kenai Peninsula, Alaska. Forest Ecology and Management 227(3):207–218.

Flint, C.G., B. McFarlane, and M. Müller. 2009. Human dimensions of forest disturbance by insects: An international synthesis. Environmental Management 43(6):1174–1186.

Francis, A. 2015. Akwesasne moves to ban genetically modified food. Aboriginal Peoples Television Network, June 29. Available at https://aptnnews.ca/2015/06/29/akwesasne-moves-ban-genetically-modified-food. Accessed November 21, 2018.

Frewer, L.J., I.A. van der Lans, A. Fischer, M.J. Reinders, D. Menozzi, X. Zhang, I. van den Berg, and K.L. Zimmermann. 2013. Public perceptions of agri-food applications of genetic modification: A systematic review and meta-analysis. Trends in Food Science and Technology 30(2):142–152.

Friedman, S.T., and G.S. Foster. 1997. Forest genetics on federal lands in the United States: Public concerns and policy responses. Canadian Journal of Forest Research 27(3):401–408.

Fuller, L., M. Marzano, A. Peace, and C.P. Quine. 2016. Public acceptance of tree health management: Results of a national survey in the UK. Environmental Science and Policy 59(1):18–25.

Gamborg, C., and P. Sandøe. 2010. Ethical considerations regarding genetically modified trees. Pp. 163–175 in Forests and Genetically Modified Trees, Y.A. El-Kassaby and J.A. Prado, eds. Rome, Italy: FAO.

Gao, C., X. Ren, A.S. Mason, H. Liu, M. Xiao, J. Li, and D. Fu. 2014. Horizontal gene transfer in plants. Functional & Integrative Genomics 14(1):23–29.

González-Ricoy, I., and A. Gosseries, eds. 2016. Institution for Future Generations. Oxford, UK: Oxford University Press.

Graff, G., and D. Zilberman. 2016. How the “IP-regulatory” complex affects incentives to develop socially beneficial products from agricultural genomics. Pp. 68–101 in The Intellectual Property-Regulatory Complex: Overcoming Barriers to Innovation in Agricultural Genomics, E. Marden, R.N. Godfrey, and R. Manion, eds. Vancouver, Canada: UBC Press.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Gugger, P.F., C.T. Liang, V.L. Sork, P. Hodgskiss, and J.W. Wright. 2018. Applying landscape genomic tools to forest management and restoration of Hawaiian koa (Acacia koa) in a changing environment. Evolutionary Applications 11(2):231–242.

Gupta, N., A. Fischer, and L. Frewer. 2012. Socio-psychological determinants of public acceptance of technologies: A review. Public Understanding of Science 21(7):782–795.

Guston, D.H. 2014. Understanding “anticipatory governance.” Social Studies of Science 44(2):218–242.

Guy, R.D. 2014. The early bud gets to warm. New Phytologist 202(1):7–9.

Hajjar, R., and R.A. Kozak. 2015. Exploring public perceptions of forest adaptation strategies in Western Canada: Implications for policy-makers. Forest Policy and Economics 61:59–69.

Hajjar, R., E. McGuigan, M. Moshofsky, and R.A. Kozak. 2014. Opinions on strategies for forest adaptation to future climate conditions in western Canada: Surveys of the general public and leaders of forest-dependent communities. Canadian Journal of Forest Research 44(12):1525–1533.

Hall, C. 2007. GM technology in forestry: Lessons from the GM food “debate.” International Journal of Biotechnology 9(5):436–447.

Hampe, A., and R.J. Petit. 2005. Conserving biodiversity under climate change: The rear edge matters. Ecology Letters 8(5):461–467.

Hamrick, J.L. 2004. Response of forest trees to global environmental changes. Forest Ecology and Management, 197(1–3):323–335.

Harry, D. 2001. Biopiracy and globalization: Indigenous peoples face a new wave of colonialism. Indigenous People Council on Biocolonialism. Available at http://www.ipcb.org/publications/other_art/globalization.html. Accessed November 21, 2018.

Harwood, C., and A. Ruuska. 2013. The personhood of trees: Living artifacts in the Upper Peninsula of Michigan. Time and Mind 3(2):135–157.

Hauser, O.P., D.G. Rand, A. Peysakhovich, and M.A. Nowak. 2014. Co-operating with the future. Nature 5111(7508):220–223.

Hettinger, N., and B. Throop. 1999. Refocusing ecocentrism: De-emphasizing stability and defending wildness. Environmental Ethics 21(1):3–21.

Higgs, E. 2005. The two culture problem: Ecological restoration and the integration of knowledge. Restoration Ecology 13(1):159–164.

Hinchee, M., C. Zhang, S. Chang, M. Cunningham, W. Hammond, and N. Nehru. 2011. Biotech Eucalyptus can sustainably address society’s need for wood: The example of freeze tolerant eucalyptus in the southeastern US. BMC Proceedings 5(Suppl 7):124.

Howe, G.T., S.N. Aitken, D.B. Neale, K.D. Jermstad, N.C. Wheeler, and T.H.H. Chen. 2003. From genotype to phenotype: Unraveling the complexities of cold adaptation in forest trees. Canadian Journal of Botany 81(12):1247–1266.

IEN (Indigenous Environmental Network). 2016. GE Trees and Indigenous Lands, February 9. Available at http://www.ienearth.org/ge-trees-and-indigenous-lands. Accessed November 21, 2018.

Jepson, P.R., and I. Arakelyan. 2017a. Developing publicly acceptable tree health policy: Public perceptions of tree-breeding solutions to ash dieback among interested publics in the UK. Forest Policy and Economics 80:167–177.

Jepson, P., and I. Arakelyan. 2017b. Exploring public perceptions of solutions to tree diseases in the UK: Implications for policy-makers. Environmental Science and Policy 76:70–77.

Johnson, D.G., and J.M. Wetmore, eds. 2009. Technology and Society: Building Our Sociotechnical Future. Cambridge, MA: MIT Press.

Johnson, L. 1991. A Morally Deep World. Cambridge, UK: Cambridge University Press.

Jolley, R., T. Punke, R. Ringeisen, L. Selzer, J.D. Williams 2017. Final Report: The Blue Ribbon Commission on Forest and Forest Products Research & Development in the 21st Century. Greenville, SC: U.S. Endowment for Forestry and Communities.

Justus, J., M. Colyvan, H. Regan, and L. Maguire. 2009. Buying into conservation: Intrinsic v. instrumental value. Trends in Ecology and Evolution 24(4):187–191.

Kahan, D., H. Jenkins-Smith, and D. Braman. 2011. Cultural cognition of scientific consensus. Journal of Risk Research 14(2):147–174.

Kalaitzandonakes, N., J.M. Alston, K.J. Bradford. 2007. Compliance costs for regulatory approval of new biotech crops. Nature Biotechnology 25(5):509–511.

Katz, E. 1992. The big lie: Human restoration of nature. Research in Philosophy and Technology 12:231–241.

Kazana, V., L. Tsourgiannis, V. Iakovoglou, C. Stamatiou, A. Alexandrov, S. Araújo, S. Bogdan, G. Božič, R. Brus, G. Bossinger, A. Boutsimea, N. Celepirović, H. Cvrčková, M. Fladung, M. Ivanković, A. Kazaklis, P. Koutsona, Z. Luthar, P. Máchová, J. Malá, K. Mara, M. Mataruga, J. Moravcikova, D. Paffetti, J. Paiva, D. Raptis, C. Sanchez, S. Sharry, T. Salaj, M. Šijačić-Nikolić, N. Tel-Zur, I. Tsvetkov, C. Vettori, and N. Vidal. 2015. Public attitudes towards the use of transgenic forest trees: A cross-country pilot survey. iForest 9:344–353.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Kazana, V., L. Tsourgiannis, V. Iakovoglou, C. Stamatiou, A. Alexandrov, S. Araújo, S. Bogdan, G. Božič, R. Brus, G. Bossinger, A. Boutsimea, N. Celepirović, H. Cvrčková, M. Fladung, M. Ivanković, A. Kazaklis, P. Koutsona, Z. Luthar, P. Máchová, J. Malá, K. Mara, M. Mataruga, J. Moravcikova, D. Paffetti, J. Paiva, D. Raptis, C. Sanchez, S. Sharry, T. Salaj, M. Šijačić-Nikolić, N. Tel-Zur, I. Tsvetkov, C. Vettori, and N. Vidal. 2016. Public knowledge and perceptions of safety issues towards the use of genetically modified forest trees: A cross-country pilot survey. Pp. 223–244 in Biosafety of Forest Transgenic Trees: Improving the Scientific Basis for Safe Tree Development and Implementation of EU Policy Directives, C. Vettori, F. Gallardo, H. Häggman, V. Kazana, F. Migliacci, G. Pilate, and M. Fladung, eds. Dordrecht, The Netherlands: Springer.

Keller, S.R., N. Levsen, M.S. Olson, and P. Tiffin. 2012. Local adaptation in the flowering-time gene network of balsam poplar, Populus balsamifera L. Molecular Biology and Evolution 29(10):3143–3152.

Kellstedt, P.M., S. Zahran, and A. Vedlitz. 2008. Personal efficacy, the information environment, and attitudes toward global warming and climate change in the United States. Risk Analysis 28(1):113–126.

Kimmerer, R. 2011. Restoration and reciprocity: The contributions of traditional ecological knowledge. Pp. 257–276 in Human Dimensions of Ecological Restoration Integrating Science, Nature, and Culture, D. Egan, E.E. Hjerpe, and J. Abrams, eds. Washington, DC: Island Press.

Kinloch, B.B., Jr., R.A. Sniezko, and G.E. Dupper. 2003. Origin and distribution of Cr2, a gene for resistance to white pine blister rust in natural populations of western white pine. Phytopathology 93(6):691–694.

Kinloch, B.B., Jr., R.A. Sniezko, and G.E. Dupper. 2004. Virulence gene distribution and dynamics of the white pine blister rust pathogen in western North America. Phytopathology 94(7):751–758.

Kooistra, C.M., and T.E. Hall. 2014. Understanding public support for forest management and economic development options after a mountain pine beetle outbreak. Journal of Forestry 112(2):221–229.

Kremer, A., O. Ronce, J.J. Robledo-Arnuncio, F. Guillaume, G. Bohrer, R. Nathan, J.R. Bridle, R. Gomulkiewicz, E.K. Klein, K. Ritland, A. Kuparinen, S. Gerber, and S. Schueler. 2012. Long-distance gene flow and adaptation of forest trees to rapid climate change. Ecology Letters 15(4):378–392.

Kunda, Z. 1990. The case for motivated reasoning. Psychological Bulletin 108(3):480–498.

Lamont, J., and C. Favor. 2017. Distributive justice. In The Stanford Encyclopedia of Philosophy (winter edition), E.N. Zalta, ed. Available at https://plato.stanford.edu/archives/win2017/entries/justice-distributive. Accessed November 21, 2018.

Leopold, A. 1949. A Sand County Almanac. Oxford, UK: Oxford University Press.

Levsen, N.D., P. Tiffin, and M.S. Olson. 2012. Pleistocene speciation in the genus Populus (Salicaceae). Systematic Biology 61(3):401–412.

Liepelt, S., R. Bialozyt, and B. Ziegenhagen. 2002. Wind-dispersed pollen mediates postglacial gene flow among refugia. Proceedings of the National Academy of Sciences of the United States of America 99(22):14590–14594.

Lombardo, L. 2014. Genetic use restriction technologies: A review. Plant Biotechnology Journal 12(8):995–1005.

Lorentz, K.A., and P.J. Minogue. 2015. Exotic Eucalyptus plantations in the southeastern US: Risk assessment, management and policy approaches. Biological Invasions 17(6):1581–1593.

Lotterhos, K.E., and M.C. Whitlock. 2015. The relative power of genome scans to detect local adaptation depends on sampling design and statistical method. Molecular Ecology 24(5):1031–1046.

Lovett, G. 2018. The risk to forest health posed by biotech-modified trees. Presentation to the National Academies of Sciences, Engineering, and Medicine’s Committee on the Potential for Biotechnology to Address Forest Health, February 8, Washington, DC.

Lucht, J.M. 2015. Public acceptance of plant biotechnology and GM crops. Viruses 7(8):4254–4281.

Lull, R.B., and D.A. Scheufele. 2017 Understanding and overcoming fear of the unnatural in discussion of GMOs. Pp. 409–419 in The Oxford Handbook of the Science of Science Communication, K.H. Jamieson, D.M. Kahan, and D.A. Scheufele, eds. New York: Oxford University Press.

Mackenzie, B.F., and B. Larson. 2010. Participation under time constraints: Landowner perceptions of rapid response to the emerald ash borer. Society and Natural Resources 23(10):1013–1022.

Manel, S., and R. Holderegger. 2013. Ten years of landscape genetics. Trends in Ecology & Evolution 28(10):614–621.

Manel, S., M.K. Schwartz, G. Luikart, and P. Taberlet. 2003. Landscape genetics: Combining landscape ecology and population genetics. Trends in Ecology & Evolution 18(4):189–197.

Manfredo, M.J., T.L. Teel, and A.D. Bright. 2004. Application of the concepts of values and attitudes in human dimensions of natural resources research. Pp. 271–282 in Society and Natural Resources: A Summary of Knowledge, M. Manfredo, J. Vaske, B. Bruyere, D. Field, and P. Brown, eds. Jefferson City, MO: Modern Litho.

McFarlane, B.L., and D. Watson. 2008. Perceptions of ecological risk associated with mountain pine beetle (Dendroctonus ponderosae) infestations in Banff and Kootenay National Parks of Canada. Risk Analysis 28(3):203–212.

McGregor, D. 2002. Indigenous knowledge in sustainable forest management: Community-based approaches achieve greater success. The Forestry Chronicle 78(6):833–836.

McLachlan, J.S., J.S. Clark, and P.S. Manos. 2005. Molecular indicators of tree migration capacity under rapid climate change. Ecology 86(8):2088–2098.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

McLane, S.C., and S.N. Aitken. 2012. Whitebark pine (Pinus albicaulis) assisted migration potential: Testing establishment north of the species range. Ecological Applications 22(1):142–153.

McManama, B.J. 2018. Indigenous perspective: Biotechnology versus the rights of Mother Earth. Webinar presentation to the National Academies of Sciences, Engineering, and Medicine’s Committee on the Potential for Biotechnology to Address Forest Health, March 2.

MEA (Millennium Ecosystem Assessment). 2005. Ecosystems and Human Well-Being: Synthesis. Washington, DC: Island Press.

Miller, D. 2017. Justice. In The Stanford Encyclopedia of Philosophy (fall edition), E.N. Zalta, ed. Available at https://plato.stanford.edu/archives/fall2017/entries/justice. Accessed November 21, 2018.

Minteer, B.A., and J.P. Collins. 2012. Species conservation, rapid environmental change, and ecological ethics. Nature Education Knowledge 3(10):14

Montwé, D., M. Isaac-Renton, A. Hamann, and H. Spiecker, 2018. Cold adaptation recorded in tree rings highlights risks associated with climate change and assisted migration. Nature Communications 9(1):1574.

Moore, K.D. 2007. In the shadow of the cedars: The spiritual values of old-growth forests. Conservation Biology 21(4):1120–1123.

Müller, M., and H. Job. 2009. Managing natural disturbance in protected areas: Tourists’ attitude towards the bark beetle in a German national park. Biological Conservation 142(2):375–383.

NASEM (National Academies of Sciences, Engineering, and Medicine). 2016a. Gene Drives on the Horizon: Advancing Science, Navigating Uncertainty, and Aligning Research with Public Values. Washington, DC: The National Academies Press.

NASEM. 2016b. Genetically Engineered Crops: Experiences and Prospects. Washington, DC: The National Academies Press.

Neale, D.B., and A. Kremer. 2011. Forest tree genomics: Growing resources and applications. Nature Reviews Genetics 12:111–122.

Needham, M., G. Howe, and J. Petit. 2016. Forest Health Biotechnologies: What Are the Drivers of Public Acceptance? Interim report and preliminary findings for the Forest Health Initiative and U.S. Endowment for Forestry and Communities. Available at https://www.foresthealthinitiative.org/resources/biotech_public_acceptance_Needham.pptx. Accessed November 21, 2018.

Nelson, C.D., W.A. Powell, S.A. Merkle, J.E. Carlson, F.V. Hebard, N. Islam-Faridi, M.E. Staton, and L. Georgi. 2014. Biotechnology of trees: Chestnut. Pp. 3–35 in Tree Biotechnology, K.G. Ramawat, J.M. Merillon, and M.R. Ahuja, eds. Boca Raton, FL: CRC Press.

Nelson, M.K., ed. 2008. Original Instructions: Indigenous Teachings for a Sustainable Future. Rochester, VT: Bear & Company.

Nelson, M., and J.B. Callicott, eds. 2008. The Wilderness Debate Rages On: Continuing the Great New Wilderness Debate. Athens, GA: University of Georgia Press.

Neumann, P.D., N.T. Krogman, and B.R. Thomas. 2007. Public perceptions of hybrid poplar plantations: Trees as an alternative crop. International Journal of Biotechnology 9(5):468–483.

Newhouse, A.E., L.D. McGuigan, K.A. Baier, K.E. Valletta, W.H. Rottmann, T.J. Tschaplinski, C.A. Maynard, and W.A. Powell. 2014. Transgenic American chestnuts show enhanced blight resistance and transmit the trait to T1 progeny. Plant Science 228:88–97.

Nilausen, C., N. Gélinas, and G. Bull. 2016. Perceived acceptability of implementing marker-assisted selection in the forests of British Columbia. Forests 7(11):286.

Nonić, M., U. Radojević, J. Milovanović, M. Perović, and M. Šijačić-Nikolić. 2015. Comparative analysis of students’ attitudes toward implementation of genetically modified trees in Serbia. iForest 8:714–718.

Norton-Smith, K., K. Lynn, K. Chief, K. Cozzetto, J. Donatuto, M.H. Redsteer, L.E. Kruger, J. Maldonado, C. Viles, and K.P. Whyte. 2016. Climate change and indigenous peoples: A synthesis of current impacts and experiences. Portland, OR: U.S. Forest Service.

Nowak, D.J., S. Hirabayashi, A. Bodine, and E. Greenfield. 2014. Tree and forest effects on air quality and human health in the United States. Environmental Pollution 193:119–129.

NRC (National Research Council). 2002. National Capacity in Forestry Research. Washington, DC: National Academy Press.

NRC. 2015. Public Engagement on Genetically Modified Organisms: When Science and Citizens Connect. Workshop Summary. Washington, DC: The National Academies Press.

Olson, M.S., N.Levsen, R.Y. Soolanayakanahally, R.D. Guy, W.R. Schroeder, S.R. Keller, and P. Tiffin. 2013. The adaptive potential of Populus balsamifera L. to phenology requirements in a warmer global climate. Molecular Ecology 22(5):1214–1230.

Orr, H.A. 2005. The genetic theory of adaptation: A brief history. Nature Reviews Genetics 6(2):119.

Oswalt, S.N., P.D. Miles, S.A. Pugh, and W.B. Smith. 2018. Forest Resources of the United States, 2017: A Technical Document Supporting the Forest Service 2020 Update of the RPA Assessment. Washington, DC: U.S. Forest Service.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Owen, R., P. Macnaghten, and J. Stilgoe. 2012. Responsible research and innovation: From science in society to science for society, with society. Science and Public Policy 39(6):751–760.

Patterson, N. 2018. Haudenosaunee & the TG chestnut. Webinar presentation to the National Academies of Sciences, Engineering, and Medicine Committee on the Potential for Biotechnology to Address Forest Health, March 2.

Pearce, D., G. Atkinson, and S. Mourato. 2006. Cost Benefit Analysis and the Environment: Recent Developments. Paris, France: Organisation for Economic Co-operation and Development.

Peterson St-Laurent, G., S. Hagerman, and R. Kozak. 2018. What risks matter? Public views about assisted migration and other climate-adaptive reforestation strategies. Climatic Change 151(3–4):573–587.

Poland, T.M., M.R. Emery, T. Ciaramitaro, E. Pigeon, and A. Pigeon. 2017. Emerald ash borer, black ash, and Native American basketmaking; Invasive insects, forest ecosystems and cultural practices. Pp. 127–140 in Biodiversity, Conservation, and Environmental Management in the Great Lakes Basin, E. Freedman and M. Meuzil, eds. Abingdon, UK: Routledge.

Porth, I., and Y.A. El Kassaby. 2014. Current status of the development of genetically modified (GM) forest trees worldwide: A comparison with the development of other GM plants in agriculture. CAB Reviews 9(008):1–12.

Poudyal, N.C., J.M. Bowker, and R.L. Moore. 2016. Understanding public knowledge and attitudes toward controlling hemlock woolly adelgid on public forests. Journal of Forestry 114(6):619–628.

Powell, W. 2014. The American chestnut’s genetic rebirth. Scientific American 310:68–73.

Powell, W. 2016. New genetically engineered American chestnut will help restore the decimated, iconic tree. The Conversation, January 19. Available at http://theconversation.com/new-genetically-engineered-american-chestnut-will-helprestore-the-decimated-iconic-tree-52191. Accessed November 21, 2018.

Pye, J.M., T.P. Holmes, J.P. Presteman, and D.N. Wear. 2011. Economic impacts of the southern pine beetle. Pp. 213–222 in Southern Pine Beetle II, R.N. Coulson, R.N. and K.D. Klepzig, eds. Asheville, NC: U.S. Forest Service.

Rask, M., and R. Worthington. 2015. Governing Biodiversity through Democratic Deliberation. New York: Routledge.

Reid, A.M., W.K. Chapman, and C.E. Prescott. 2016. Comparing lodgepole pine growth and disease occurrence at six Long-Term Soil Productivity (LTSP) sites in British Columbia, Canada. Canadian Journal of Forest Research 46(4):595–599.

Reo, N., and L. Ogden. 2018. Anishnaabe Aki: An indigenous perspective on the global threat of invasive species. Sustainability Science 13(5):1443–1452.

Rice, K.J., D.R. Gordon, J.L. Hardison, and J.M. Welker. 1993. Phenotypic variation in seedlings of a “keystone” tree species (Quercus douglasii): The interactive effects of acorn source and competitive environment. Oecologia, 96(4):537–547.

Richardson, A.O., and J.D. Palmer. 2007. Horizontal gene transfer in plants. Journal of Experimental Botany 58(1):1–9.

Roberts, R.M. 2005. Walking backwards into the future: Māori views on genetically modified organisms. Perspectives on Indigenous Knowledge, WINHEC Journal. Available at http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.562.3006&rep=rep1&type=pdf. Accessed November 21, 2018.

Rokeach, M. 1973. The Nature of Human Values. New York: Free Press.

Rolston, H. 1988. Environmental Ethics: Duties to and Values in the Natural World. Philadelphia, PA: Temple University Press.

Rolston, H. 1998. Aesthetic experience in forests. Journal of Aesthetics and Art Criticism 56(2):157–166.

Rosenberger, R.S., E.M. White, J.D. Kline, and C. Cvitanovich. 2017. Recreation Economic Values for Estimating Outdoor Recreation Economic Benefits from the National Forest System. Portland, OR: U.S. Forest Service.

Sandler, R. 2012. Intrinsic value, ecology, and conservation. Nature Education Knowledge 3(10):4.

Sandler, R. 2013. Climate change and ecosystem management. Ethics, Policy and Environment 16(1):1–15.

Sandler, R. 2018. Forest biotechnology: Environmental ethics perspectives. Webinar presentation to the National Academies of Sciences, Engineering, and Medicine’s Committee on the Potential for Biotechnology to Address Forest Health, March 12.

Sarewitz, D. 2011. Anticipatory governance of emerging technologies. Pp. 95–105 in The Growing Gap Between Emerging Technologies and Legal-Ethical Oversight, G.E. Marchant, B.R. Allenby, and J.R. Herkert, eds. Dordrecht, The Netherlands: Springer.

Savolainen, O., and T. Pyhäjärvi. 2007. Genomic diversity in forest trees. Current Opinion in Plant Biology 10(2):162–167.

Savolainen, O., T. Pyhäjärvi, and T. Knürr. 2007. Gene flow and local adaptation in trees. Annual Review of Ecology, Evolution, and Systematics 38:595–619.

Scheufele, D.A. 2006. Messages and heuristics: How audiences form attitudes about emerging technologies. Pp. 20–25 in Engaging Science: Thoughts, Deeds, Analysis and Action, J. Turney, ed. London, UK: The Wellcome Trust.

Schmidtling, R.C. 2003. The southern pines during the Pleistocene. Acta Horticulturae 615:203–209.

Sedjo, R.A. 2006. Toward Commercialization of Genetically Engineered Forests: Economic and Social Considerations. Washington, DC: Resources for the Future.

Sedjo, R.A. 2010. Social, legal and regulatory issues related to transgenic trees. Pp. 203–226 in Forests and Genetically Modified Trees, Y.A. El-Kassaby and J.A. Prado, eds. Rome, Italy: FAO.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

Selin, C., K.C. Rawlings, K. de Ridder-Vignone, J. Sadowski, C. Altamirano Allende, G. Gano, S.R. Davies, and D.H. Guston. 2017. Experiments in engagement: Designing public engagement with science and technology for capacity building. Public Understanding of Science 26(6):634–649.

Semizer-Cuming, D., E.D. Kjær, and R. Finkeldey. 2017. Gene flow of common ash (Fraxinus excelsior L.) in a fragmented landscape. PLoS One 12(10):e0186757.

SFI (Sustainable Forestry Initiative). 2015. SFI policies. Section 7 in SFI 2015–2019 Standards and Rules. Available at http://www.sfiprogram.org/files/pdf/2015-2019-standardsandrules-web-lr-pdf. Accessed November 23, 2018.

Shafer, A., C.I. Cullingham, S.D. Cote, and D.W. Coltman. 2010. Of glaciers and refugia: A decade of study sheds new light on the phylogeography of northwestern North America. Molecular Ecology 19(21):4589–4621.

Siipi, H. 2015. Is genetically modified food unnatural? Journal of Agricultural and Environmental Ethics 28(5):807–816.

Sikes, P. 2006. Decolonizing research and methodologies: Indigenous peoples and cross-cultural contexts. Pedagogy, Culture & Society 14(3):349–358.

Sjöberg, L. 2004. Principles of risk perception applied to gene technology. EMBO Reports 5:S47–S51.

Slovic, P. 2000. The Perception of Risk. London, UK: Earthscan.

Slovic, P. 2010. The Feeling of Risk: New Perspectives on Risk Perception. New York: Earthscan.

Smith, I. 2016. The Intrinsic Value of Endangered Species. New York: Routledge.

Smith, L.T. 2013. Decolonizing Methodologies: Research and Indigenous Peoples. London, UK: Zed Books Ltd.

Smolker, R. 2018. Biotechnology and forest health. Webinar presentation to the National Academies of Sciences, Engineering, and Medicine Committee on the Potential for Biotechnology to Address Forest Health, March 27.

Sniezko, R.A., and J. Koch. 2017. Breeding trees resistant to insects and diseases: Putting theory into application. Biological Invasions 19(11):3377–3400.

Soolanayakanahally, R.Y., R.G. Guy, S.N. Silim, E.C. Drewes, and W.R. Schroeder. 2009. Enhanced assimilation rate and water use efficiency with latitude through increased photosynthetic capacity and internal conductance in balsam poplar (Populus balsamifera L.). Plant, Cell & Environment 32(12):1821–1832.

Sork, V.L., S.N. Aitken, R.J. Dyer, A.J. Eckert, P. Legendre, and D.B. Neale. 2013. Putting the landscape into the genomics of trees: Approaches for understanding local adaptation and population responses to changing climate. Tree Genetics & Genomes 9(4):901–911.

Soulé, M.E. 1985. What is conservation biology? Bioscience 35(11):727–734.

Stilgoe, J., R. Owen, and P. Macnaghten, P. 2013. Developing a framework for responsible innovation. Research Policy 42(9):1568–1580.

Strauss, S.H., M. Schmitt, and R. Sedjo. 2009. Forest scientist views of regulatory obstacles to research and development of transgenic forest biotechnology. Journal of Forestry 107(7):350–357.

Strauss, S.H., K.N. Jones, H. Lu, J.D. Petit, A.L. Klocko, M.G. Betts, B.J. Brosi, R.J. Fletcher Jr., and M.D. Needham. 2017. Reproductive modification in forest plantations: Impacts on biodiversity and society. New Phytologist 213(3):1000–1021.

Thompson, D. 2010. Representing future generations: Political presentism and democratic trusteeship, Critical Review of International Social and Political Philosophy, 13(1):17–37.

Tomback, D., and P. Achuff. 2010. Blister rust and western forest biodiversity: Ecology, values and outlook for white pines. Forest Pathology 40(3–4):186–225.

Tomblin, D., Z. Pirtle, M. Farooque, D. Sittenfeld, E. Mahoney, R. Worthington, and D. Guston. 2017. Integrating public deliberation into engineering systems: Participatory technology assessment of NASA’s asteroid redirect mission. Astropolitics 15(2):141–166.

Tsourgiannis, L., V. Kazana, A. Karasavvoglou, M. Nikolaidis, G. Florou, and P. Polychronidou. 2013. Exploring consumers’ attitudes towards wood products that could be derived from transgenic plantations in Greece. Procedia Technology 8:554–560.

Tsourgiannis, L., V. Kazana, and V. Iakovoglou. 2015. Exploring the potential behavior of consumers towards transgenic forest products: The Greek experience. iForest 8:707–713.

Tsourgiannis, L., V. Kazana, and V. Iakovoglou. 2016. A comparative analysis of consumers’ potential purchasing behaviour towards transgenic-derived forest products: The Greek case. Pp. 245–260 in Biosafety of Forest Transgenic Trees: Improving the Scientific Basis for Safe Tree Development and Implementation of EU Policy Directives, C. Vettori, F. Gallardo, H. Häggman, V. Kazana, F. Migliacci, G. Pilate, and M. Fladung, eds. Dordrecht, The Netherlands: Springer.

UN (United Nations). 1992. Convention on Biological Diversity. Available at https://www.cbd.int/convention/text/default.shtml. Accessed November 23, 2018.

UN. 2006. Declaration on the Rights of Indigenous Peoples. Available at https://www.ohchr.org/en/issues/ipeoples/pages/declaration.aspx. Accessed November 24, 2018.

U.S. Census Bureau. 2017. North American Industry Classification System. Washington, DC: U.S. Office of Management and Budget.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×

USDA-FS (U.S. Department of Agriculture’s Forest Service). 2007. Valuing Ecosystem Services: Capturing the True Value of Nature’s Capital. Available at https://www.fs.fed.us/ecosystemservices/pdf/ecosystem-services.pdf. Accessed November 23, 2018.

USDA-FS. 2018. Forest Health Highlights in Washington—2017. Available at https://www.dnr.wa.gov/publications/rp_fh_2017_forest_health_highlights.pdf. Accessed December 6, 2018.

Van Deynze, A., K.J. Bradford, H. Daniell, J.M. DiTomaso, N. Kalaitzandonakes, C. Mallory-Smith, C.N. Stewart, S.H. Strauss, and R. Van Acker. 2016. The science of gene flow in agriculture and its role in coexistence. Pp. 13–37 in The Coexistence of Genetically Modified, Organic and Conventional Foods, N. Kalaitzandonakes, P.W.B. Phillips, J. Wesseler, and S.J. Smyth, eds. New York: Springer.

Vaske, J.J., and M.P. Donnelly. 1999. A value-attitude-behavior model predicting wildland preservation voting intentions. Society and Natural Resources 12(6):523–537.

Vaske, J.J., and M.J. Manfredo. 2012. Social psychological considerations in wildlife management. Pp. 43–57 in Human Dimensions of Wildlife Management, D.J. Decker, S. Riley, and W.F. Siemer, eds. Baltimore, MD: Johns Hopkins University Press.

Vaske, J.J., and D. Whittaker. 2004. Normative approaches to natural resources. Pp. 283–294 in Society and Natural Resources: A Summary of Knowledge, M. Manfredo, J. Vaske, B. Bruyere, D. Field, and P. Brown, eds. Jefferson City, MO: Modern Litho.

Warwell, M.V. 2015. Genecology and Phenotypic Selection in Whitebark Pine (Pinus albicaulis) and Ponderosa Pine (Pinus ponderosa) Under Warm-Dry Climate. Doctor of Philosophy (Ph.D.) Dissertation, University of Minnesota. Available at https://conservancy.umn.edu/bitstream/handle/11299/180219/Warwell_umn_0130E_16397.pdf?sequence=1&isAllowed=y. Accessed December 6, 2018.

Warwell, M.V., and R.G. Shaw. 2017. Climate-related genetic variation in a threatened tree species, Pinus albicaulis. American Journal of Botany 104(8):1205–1218.

Warwell, M.V., and R.G. Shaw. 2018. Phenotypic selection on growth rhythm in whitebark pine under climatic conditions warmer than seed origins. Journal of Evolutionary Biology 31(9):1284–1299.

Whyte, K.P., M. Dockry, W. Baule, and D. Fellman. 2014. Supporting tribal climate change adaptation planning through community participatory strategic foresight scenario development. In Project Reports, D. Brown, W. Baule, L. Briley, and E. Gibbons, eds. Great Lakes Integrated Sciences and Assessments (GLISA) Center. Available at http://glisa.umich.edu/media/files/projectreports/GLISA_ProjRep_Strategic-Foresight.pdf. Accessed November 17, 2018.

Winner, L. 2010. The Whale and the Reactor: A Search for Limits in an Age of High Technology. Chicago, IL: University of Chicago Press.

Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 93
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 94
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 95
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 96
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 97
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 98
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 99
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 100
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 101
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 102
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 103
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 104
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 105
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 106
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 107
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 108
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 109
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 110
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 111
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 112
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 113
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 114
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 115
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 116
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 117
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 118
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 119
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 120
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 121
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 122
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 123
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 124
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 125
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 126
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 127
Suggested Citation:"4 Ecological, Economic, Social, and Ethical Considerations in the Use of Biotechnology in Forest Trees." National Academies of Sciences, Engineering, and Medicine. 2019. Forest Health and Biotechnology: Possibilities and Considerations. Washington, DC: The National Academies Press. doi: 10.17226/25221.
×
Page 128
Next: 5 An Integrated Impact Assessment Framework »
Forest Health and Biotechnology: Possibilities and Considerations Get This Book
×
Buy Paperback | $90.00 Buy Ebook | $69.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

The American chestnut, whitebark pine, and several species of ash in the eastern United States are just a few of the North American tree species that have been functionally lost or are in jeopardy of being lost due to outbreaks of pathogens and insect pests. New pressures in this century are putting even more trees at risk. Expanded human mobility and global trade are providing pathways for the introduction of nonnative pests for which native tree species may lack resistance. At the same time, climate change is extending the geographic range of both native and nonnative pest species.

Biotechnology has the potential to help mitigate threats to North American forests from insects and pathogens through the introduction of pest-resistant traits to forest trees. However, challenges remain: the genetic mechanisms that underlie trees' resistance to pests are poorly understood; the complexity of tree genomes makes incorporating genetic changes a slow and difficult task; and there is a lack of information on the effects of releasing new genotypes into the environment.

Forest Health and Biotechnology examines the potential use of biotechnology for mitigating threats to forest tree health and identifies the ecological, economic, and social implications of deploying biotechnology in forests. This report also develops a research agenda to address knowledge gaps about the application of the technology.

  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. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

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

    « Back Next »
  7. ×

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

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    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!