This is a good article. Click here for more information.

Assisted migration of forests in North America

From Wikipedia, the free encyclopedia

The western larch (Larix occidentalis) was selected by the government of British Columbia for an assisted migration program that saw the tree being seeded 1000 km north of its historical range.

The assisted migration of forests in North America is an ongoing process of human-facilitated forest migration applied to a variety of tree species on the continent. Programs created by public and indigenous governmental bodies, private forest owners, and land trusts have been researching, testing, evaluating, and sometimes implementing forest assisted migration projects as a form of adaptation to climate change.[1][2] Assisted migration in the forestry context differs from assisted migration as originally proposed in the context of conservation biology, where it is regarded as a management tool for helping endangered species cope with the need for climate adaptation.[3] The focus in forestry is mitigating climate change's negative effects on the health and productivity of working forests.

Forestry assisted migration is already underway in North America because of the rapidly changing climate and the forestry industry's reseeding practices. It is now standard practice for governmental and industrial harvests of trees to be followed by the planting of seeds or seedlings in the harvested areas. Hence, an opportunity automatically arises post-harvest to select seeds (and sometimes different species of trees) from areas with climates that are expected to arrive in the harvested sites decades into the future. The government of British Columbia in Canada was the first federated state on the continent to make the decision to change their seed transfer guidelines accordingly in 2009.[4][5][6]

Longer distance forms of assisted migration were not, however, considered prior to climate modelling and within-forest evidence of the increasing pace of climate change. Serious discussion and debate ensued in the forestry profession beginning around 2008.[7] The debate focuses around the ethical implications of artificially migrating ecosystems, the risks and benefits of such endeavors, and the values at the heart of assisted forest migration projects.[8]

There are also recorded instances of inadvertent assisted migration of North American trees. Beginning in the early 20th century, two trees famously endemic to California, the giant sequoia and coast redwood, have been planted for urban forestry purposes northward in cities along the Pacific coast of Oregon, Washington, and British Columbia. Today these specimens are not only thriving; they are prominent along urban skylines and often outrank the native trees in sizes achieved. As well, several kinds of Magnolia native to the southeastern United States have dispersed into the forest understory, thanks to ornamental plantings producing seeds beyond their native ranges.

Background[]

Change in photosynthetic activity in northern forests from 1982 to 2003[9]

Climate change is increasing the average temperatures of North American forests. In the contiguous United States forests overall have experienced a 0.8 °C increase since 1900, and the Canadian boreal forest increased by 1.5 °C in the same period.[10] There is evidence that some areas within the western boreal forest in British Columbia have increased by 2 °C since 1948, and there is a high likelihood that these regions will rise by another 3 or 4 degrees before the end of the century, thus permanently changing this ecosystem.[11] As well, the health of the boreal forest is threatened not only by climate heating but also drying, which compounds problematic shifts in this northern-most forest type.[12][13]

Changes that happen too rapidly will exceed the ability of forest plants to keep pace by shifting their ranges. Range shifts upslope to higher topographic altitudes require less distance than shifts tracking climate over relatively flat terrain poleward.[14][7][15] Species within the southern limits of their range are already showing decline or extirpation, including some of North America's most iconic trees (Joshua Tree and giant sequoia).[16] Drier conditions in the American Southwest are leading to shrubby grassland replacing pinyon-juniper woodland, following drought-induced die-off of trees.[17] Foresters in Minnesota are predicting that prairie ecosystems will expand, replacing forests as their region becomes warmer and drier.[18] Regional climate change in the Sierra mountains of California made it possible for native bark beetles to kill drought-stressed conifer trees, which in turn led to expansive and catastrophic fires.[19] Although the giant sequoia is more resistant to bark beetles than are the pines and firs intermixed in and surrounding its groves, the regional fires consuming the standing dead conifers in 2020 and 2021 proved fatal to giant sequoias whose living canopies were also engulfed.[20][21]

Definition[]

A 2016 review article defines forestry assisted migration as "the physical realignment of natural populations to the climate for which they are adapted, by reforestation in sites where their suitable climate is predicted to occur in the future, as an active management option with the aim of maintaining healthy tree ecosystems in the future."[22] The article says that human assistance in helping trees migrate is necessary because "geographic shifts of tree populations will have to be 10 to 100 times faster than they have been in the past or are at present."[22]

Three types of assisted migration in forestry[]

Three types of forestry assisted migration.[23]

There are three types of forestry assisted migration, which Natural Resources Canada describes in this way:[1]

  1. Assisted population migration: the human-assisted movement of populations within a species' established range.
  2. Assisted range expansion: the human-assisted movement of species to areas just outside their established range, facilitating or mimicking natural range expansion.
  3. Assisted long-distance migration: the human-assisted movement of species to areas far outside their established range (beyond areas accessible through natural dispersal).

The United States Forest Service uses the same three types, but sometimes refers to the third type as assisted species migration.[23] In 2018 the U.S. Forest Service published a longer policy document, which listed and defined the three types this way:[24]

  1. Assisted population migration (also assisted genetic migration or assisted gene flow[25]) – moving seed sources or populations to new locations within the historical species range.
  2. Assisted range expansion – moving seed sources or populations from their current range to suitable areas just beyond the historical species range, facilitating or mimicking natural dispersal.
  3. Assisted species migration (also species rescue, managed relocation, or assisted long-distance migration) – moving seed sources or populations to a location far outside the historical species range, beyond locations accessible by natural dispersal.

Terminology to distinguish forestry applications of assisted migration from the more controversial practices being debated within conservation biology were suggested in a lengthy 2012 report on climate adaptation within the U.S. Forest Service. The authors proposed "forestry assisted migration" for their agency's endeavors and "rescue assisted migration" for the species-specific extinction concerns of conservation biologists.[26] That same year nine forestry scientists (all but one of whom is Canadian) coauthored a paper that likewise recommended "forest assisted migration" as preferred terminology for helping their profession stay clear of the controversy that they referred to as "species rescue assisted migration."[27] As of 2021, the original classification names of the three types is in public as well as professional use, with foresters focusing on the two most moderate forms.[21]

Early scholarship and debate[]

Beginning in 2004 and accelerating in 2007, researchers in conservation biology published papers on the pros and cons of supplementing traditional management practices for preventing plant and animal extinctions with species translocation tactics to accommodate the range shifts already becoming evident as a result of climate change.[28][29][30][31] Conservation biologists also published arguments for and against the three main terms used to name the same management tool: assisted migration, assisted colonization, and managed relocation.[32][33]

Native bark beetles are now able to kill even the highest elevation trees (Engelmann spruce) in the national forests of Colorado. Photo June 2014, Wolf Creek Pass.

The focus among forestry professionals and researchers was different. Paleoecologists had already concluded that there were significant lags in northward movement of even the dominant canopy trees in North America during the thousands of years since the final glacial retreat.[34][35][36][37] In the 1990s, forestry researchers had begun applying climate change projections to their own tree species distribution models, and some results on the probable distances of future range shifts prompted attention.[7] As well, translocation terminology was not controversial among forestry researchers because "migration" was the standard term used in paleoecology for natural movements of tree species recorded in the geological record. Discussion in forestry journals therefore pertained more to questions of how and when and for which species climate-adaptive plantings and range expansions should begin.

Debate as to the need for and ethics of assisted migration was present among forestry researchers. But compared to the debate among conservation biologists,[38] it was muted and short-lived.[27][23] One of the strongest statements urging caution that appeared in a forestry journal was published in 2011. Isabelle Aubin and colleagues stated that "assisted migration is being proposed to reduce the impacts of human-induced climate change, an unprecedented situation in human history that brings with it entirely new environmental, societal and ethical challenges."[8]

But also in 2011, researchers within the Canadian Forest Service expressly distinguished forestry assisted migration from the concept debated in conservation biology.[39][40] Preventing species extinctions was not a forestry focus. Rather, forestry forms of assisted migration would be undertaken for the purposes of maintaining the forests' ecosystem services as well as the extractive resource values of timber and pulp production.[27] The Canadian Forest Service also produced research showing forest health and productivity could benefit from relocating forest understory plant species along with the dominant canopy tree species and that this would help with the successful establishment of a forest at a new location.[39]

Defusing the controversy[]

"Transition" results from assisted migration of more southerly tree species.[41]

In focusing on common canopy trees, forestry forms of assisted migration generated less intense debate than did the largely animal focus of conservation biologists.[42] Animals have the ability to move at will and thus pose risks of quickly shifting beyond or entirely out of the locales into which they are translocated. Animal mobility also complicates monitoring results, and may require the added costs of radio collars and skilled trackers. Costs of translocating mammals, in particular, may escalate owing to standards of minimizing psychological and physical suffering during capture, transport, and release.[43]

In contrast, assisted migration of forests can be done at little cost, especially when it is paired with an existing reforestation program. Risks are also reduced in forestry because the only motility that trees have occurs when their seeds are released. Tall woody plants tend to have long generation times, so several decades may pass before translocated seeds or seedlings can produce a next generation of seeds. Even then, seed dispersal distances may be limited, except for tufted seeds carried by wind and small fruits, such as berries, swallowed whole by birds and wide-ranging mammals.[44]

Communication strategies have also defused the controversy. Forest researchers and managers have talked and written about climate adaptation projects without using terminology coined by conservation biologists, whose focus is generally the wellbeing of single species of animals and plants that might be harmed or lost as climate change continues.[45][46]

By 2014 forestry managers in Canada had also honed their public communications in a number of ways, such as avoiding to be associated with the scientific debate on assisted migration and instead presenting their assisted migration as forest management best practices.[47][48]

A list of 15 tree species native to more southerly regions of the US, but which were already (or would soon become during climate change) adapted for living in the Chicago area.[49]

Communication in both word and images has also been honed in the United States, thanks in part to the Northern Institute of Applied Climate Science (NIACS), a collaborative group launched in 2010 by the U.S. Forest Service.[50] The NIACS proposed various ways though which foresters can adapt to climate change by changing their forest management techniques.[51]

Indigenous people's perspectives and actions[]

The term "assisted colonization", used in the guidelines of the International Union for Conservation of Nature (IUCN) to describe moving a species outside its native range to prevent it from going extinct, has been criticized as potentially offensive to indigenous peoples because the word "colonisation" is linked to the genocide of indigenous people.[52] According to Connie Barlow, the term 'managed relocation' may also be offensive in the United States, owing to the harm caused by the Indian Relocation Act of 1956.[53] Some scholars have pointed out that this terminology could prevent these nations from consenting to and participating in forestry assisted migration projects.[54]

Indigenous botanist Robin Wall Kimmerer has used the term "helping forests walk" in lieu of any of the terms currently used by foresters, conservation biologists, and other researchers and scholars.[55] Kimmerer is founder and director of the Center for Native Peoples and the Environment, hosted by SUNY College of Environmental Science and Forestry, where she writes, "The validity of using Traditional Ecological Knowledge as a partner to ecological science in education and research is gaining traction through our efforts. The successful development of the Center has created a platform from which grant proposals such as NSF-IGERT "Helping Forests Walk" have developed."[56] This project is also listed on the "tribal nations" page of the U.S. government "Climate Resilience Toolkit" website.[57]

Native Americans may have helped pawpaw (Asimina triloba) disperse by carrying its fruit northward as glaciers retreated in eastern North America.

The Northern Institute of Applied Climate Science facilitated collaboration on a tribal climate adaptation menu for the Great Lakes region. The document that resulted was published in 2019 as Dibaginjigaadeg Anishinaabe Ezhitwaad: A Tribal Climate Adaptation Menu. The only standard term used was "assisted migration" (and this was used only once). The term "invasive species" was replaced either by a new term, "non-local beings," or by an Ojibwe phrase, "Bakaan ingoji ga-ondaadag," which is defined as "that which comes from somewhere else and now resides here."[58] The document summarized the importance of word choice in this way:

"As the original and current stewards of the Great Lakes region, Ojibwe and Menominee tribal members who worked on this project felt it important to bring a language of parity between human and non-human beings. English and scientific terminology used in currently accepted land management practices tends to assume human dominance over non-human beings. This approach deviates from an equitable co-existence with our environment, which is typically a foundational understanding in many indigenous cultures. The terms used throughout this document are an attempt to recognize agency and sovereignty of our non-human relations."[58]

As of 2021, the Little Traverse Bay Bands of Odawa Indians in northern Michigan have planted tree species common to more southerly ranges, including shagbark hickory, silver maple, black walnut, swamp white oak, sassafras, and pawpaw. Noah Jansen, conservationist staff, explains, "I don't know which of these species are going to thrive in 50 or 100 years. So we cast the net broad and try to have something there that creates habitat for wildlife, sources of cultural significance for tribal members and areas to hunt and gather."[59]

Starting in 1492, the arrival of colonizers pushed Indigenous peoples of the Americas out of their native lands and severely constricted where they can continue to live.[60] Because California is already experiencing adverse climate change effects on native vegetation, the need for adaptive responses is severe. Traditional use of fire in managing local ecosystems for safety and for ensuring abundance of culturally significant plant and animal foods has drawn media attention to the climate-adaptive actions by the Karuk Tribe along the Klamath River of northern California.[61] But assisted migration is not a readily palatable option for Indigenous peoples, nor for their mutual relations with the ecosystems in which their cultures are enmeshed.[62]

Implementation of assisted migration in North American forests[]

Changes in seed transfer guidelines[]

Tunnels created by the larvae of native bark beetles are evident beneath the bark of this dead Douglas-fir trunk in Colorado.

Forestry research has long been conducted by governmental agencies responsible for management of federal and provincial forests in Canada (Crown lands) and federal and state forests in the United States. Because timber and pulp resources are harvested from multiple use regions of public forests in North America, re-seeding or planting seedlings onto harvested sites has been a routine management practice for many decades. Research for improving silviculture practices has included tree provenance trials, by which a variety of seed sources (or provenances) are planted together at several distinct geographic test sites ("common gardens") across the current and, increasingly, potential range of a species.[63] Because provenance trial data are available for many widespread, commercially valuable tree species, forestry professionals already have long-term experiments underway for testing tree species and population viability and performance in locations outside of native geographic ranges. Provenance trial locations poleward of native ranges or at higher (and thus cooler) elevations help forest managers determine whether, where, and when assisted migration as a climate adaptation measure should be implemented.[24][5]

In 2009, British Columbia altered its standards for selecting seeds for replanting forests after a timber harvest.[64] Previously, foresters were required to use seeds from within 300 meters downhill and 200 meters uphill, but the new policy allowed foresters to obtain seeds from up to 500 meters downhill for most species, taking advantage of the fact that populations in warmer habitats downhill may be better adapted to the future climate of the restoration site.[65] As of 2022, research, including study of provenance trials already in place, is ongoing in British Columbia[66] and the Canadian provinces of Alberta[67] and Ontario.[68] These changes were implemented partly because Canadian policymakers feared that, if they did not set the guidelines, the private sector would be tempted to pursue an unregulated assisted migration strategy on its own.[69] In 2022, the climate-adapted seed transfer guidelines for British Columbia will transition from being optional for provincial reforestation projects to becoming mandatory.[70]

Within the United States Forest Service, regional geneticists have recommended a "no regrets" approach to considering assisted migration and seed transfer as a climate adaptation strategy.[71][72] Population transfers to match seed sources to projected future conditions are recommended only for species where experience or research has demonstrated appropriate climate transfer limits.[24] Forest Service researchers have also used computer modeling to offer projections of native tree species range shifts under a variety of climate change projections. A total of 76 species of trees native to the western USA have range-shift projection maps available online.[73][74] A total of 134 species of trees native to the eastern USA have range shift projection maps available online.[75] Forest Service researchers have also been publishing regional range shift projections for North American tree species since the 1990s.[76][77]

Assisted migration of the western larch[]

Western larch in autumn

In 2010, the Government of British Columbia implemented an assisted range expansion project for one particular canopy species: the western larch. Larch could now be selected for provincial reforestation projects nearly 1,000 kilometers northward of its current range.[48] Research had shown that the western larch, the most productive of the three species of larch native to North America,[78] has no trouble growing in northern British Columbia, where climatic conditions are predicted to match the western larch's historical range by 2030.[69] This was the first government-approved assisted long-distance migration program for a North American canopy tree. The western larch was selected for assisted species migration because of its significant commercial importance and the fear that climate change and parasites such as the mountain pine beetle would considerably diminish its supply.[79]

Foresters in the United States have also initiated "experimental treatments" of larch-dominated national forests in Montana.[80] However, if some "aggressively warming climate scenarios" unfold, foresters will need to let go of any expectations of helping this species maintain a presence south of the Canadian border.[78]

Assisted migration of Whitebark pine[]

An old whitebark pine in Oregon

University of British Columbia forestry researchers Sierra McLane and Sally Aitken were among the first scientists to engage in long-distance experimental plantings to test how far northward seeds of a tree native to North America could germinate and continue to grow, in advance of expected warming in North America.[15] They selected Whitebark pine for their case study, as it is a keystone species that is already a threatened species in western North America. The authors used species distribution modelling to learn that this species will likely be extirpated from most of its current range as temperatures rise over the next half century. The same models indicate that a large area within northwestern British Columbia is already climatically suitable for the species under current conditions and will remain so throughout the 21st century.

Experimental plantings in eight sites began in 2007. Ten years later results were tallied. Protective snow cover throughout the winter, followed by spring melt ending in April or May, proved to be a far better indicator of seedling survival and growth than latitude. Indeed, the tallest seedlings measured in 2017 were those at the northern-most experimental site — 600 kilometers beyond the species' current distribution. The authors explained that "post-glacial migration has been too slow" for this large-seeded pine to track warming, even during the thousands of years that preceded the pace and scale of human-caused climate change.[81]These experimental plantings for learning the poleward limits of tolerance, reproduction, and thrival of whitebark pine are first steps toward implementation.[82]

Assisted migration of Mexico's oyamel fir[]

Oyamel fir foliage with monarch butterflies Danaus plexippus

The Monarch Butterfly Biosphere Reserve in central Mexico depends upon the integrity of its evergreen forest trees to serve as winter habitat for a long-distance annual migrator: the monarch butterfly. The oyamel fir is a major species of evergreen on which the overwintering butterflies spend a significant time during their winter diapause, or suspended development.[83] The tree's survival is threatened at its lower elevations on mountain slopes, in part, by climate change. Climate stress is also indicated by weak seedling recruitment, meaning that most of the oyamel fir seedlings don't survive past that point. This is true even in the higher forest elevations where trees do not otherwise show strong indicators of stress.[84] Upslope assisted migration experiments are underway, with findings suggesting that "400 meters upward in elevation (i.e., assisted migration) to compensate for future warmer climates does not appear to have any negative impacts on the seedlings, while potentially conferring closer alignment to future climates."[85]

Assisted migration of oyamel fir is complicated by the necessity of planting this shade-tolerant species under nurse plants — especially because of the extreme solar radiation and the large differences between day and night temperatures at high elevations.[81][86] Thus, where upslope locales are devoid of forest shade, nurse plants (e.g. Baccharis conferta) must be established first.[84]

Recent developments (2019-present)[]

This incense cedar, whose native range is in Oregon and northern California, was one of the more southerly conifer seedlings planted after a logging operation on private land on Whidbey Island in Washington state 2019.

By 2021 the projects for climate adaptation in forestry had expanded to include planting "novel species" from southward latitudes of the same continent.[87][88] An early example was the updated management plan for the Petawawa Research Forest northwest of Ottawa in the province of Ontario, Canada.[89] In addition to planting seeds from more southerly populations of the existing species in the forest, the management guidelines also authorized experiments in planting tree species whose northernmost ranges were far to the south, notably several species of hickory, Virginia pine, and American chestnut. In 2019 a small land trust in northern Michigan, the Leelanau Conservancy, began advising land owners to plant "trees whose native ranges end just south of here, yet are projected to do well in our region."[90] In Minnesota, The Nature Conservancy is engaged in a similar project of planting more southerly tree species in an attempt to maintain a fully forested north woods, even as the dominant boreal trees decline in health and numbers.[91]

In early spring 2020 a multi-group effort[92] began restoring 154 acres of a 2012 clearcut of Douglas-fir forest at a low elevation site east of Seattle, Washington. The Stossel Creek project[93] is primarily a watershed restoration effort using native tree species, but it is also experimenting with long-distance assisted population migration of the dominant canopy tree, Douglas-fir. Seedlings of Douglas-fir were sourced from nurseries several hundred miles to the south, in Oregon and even northern California. Because nursery stock had run short of native western red-cedar, project leaders chose to substitute a closely related conifer sourced from Oregon and northern California: incense cedar. This species is more drought-tolerant than western red-cedar, but all wild populations occur southward of Washington state. Including incense cedar in the planting list thus brings an element of assisted species migration into this forest restoration project. First-year survival was strong for southerly sourced seedlings of both Douglas-fir and incense cedar.[94]

Inadvertent assisted migration of the coast redwood and the giant sequoia[]

Coast redwood planted in 1948 in Seabeck, Washington along the eastern seacoast of the Olympic Peninsula.

Mature horticultural plantings of trees northward of their native ranges are a form of assisted migration experiment already underway.[95] Because the original plantings likely did not include the goal of helping the trees migrate northward in a warming climate, this form of assisted migration can be called inadvertent, or unintended. Jesse Bellemare and colleagues may have coined the term in a paper published in 2015: "It appears that a subset of native plants, particularly those with ornamental value, might already have had opportunities to shift their ranges northward via inadvertent human assistance."[96] Bellemare suggests, "Native plant horticulture is giving us some fascinating insights into what is likely to happen with climate change."[97]

This is the largest giant sequoia (Sequoiadendron giganteum) among more than a dozen that were planted early in the 20th century in Laurelhurst Park of Portland, Oregon.

The tallest tree in the world is the coast redwood; the most massive tree is the closely related giant sequoia. Both have non-overlapping native ranges limited to California, although the coast redwood extends a few miles northward along the southern-most coast of Oregon. Both species have been planted as landscaping trees in urban areas hundreds of miles northward, in latitude, of their native ranges,[98][99] including Washington state.[100]

Video-documentation and analysis of a sampling of horticultural plantings of both species of California trees reveals strong growth from Portland to Seattle — substantially north of their native ranges. While horticulturally planted coast redwood and giant sequoia regularly produce cones in this northward region, only coast redwood is documented as having fully naturalized. This is evidenced by seedlings and saplings growing nearby the original plantings.[101]

Diminishing fog along California's northern coast is already substantial,[102] and its consequences would make coast redwood increasingly "drought stressed under a summer climate of reduced fog frequency and greater evaporative demand."[103] Researchers have begun studying how diminishment of fog, owing to climate change, would make habitat unsuitable for Coast redwood in southern portions of their range.[104] Even in northern sections of current range, climate change could constrict habitability such that low slopes along rivers would become the final refugia.[105] Authors of the guest editorial in a 2021 issue of Journal of Ecology featured coast redwood in a concluding statement: "We expect that dominant tree species threatened within their range (e.g. coast redwood) would have to be translocated at a landscape level to protect overall habitat, ecosystem productivity and associated species."[106]

The sudden deaths of some previously healthy giant sequoias, which are native to the western slope of the Sierra Nevada mountain range, are correlated with California's record drought of 2012–2016. The deaths caught scientists and national park managers by surprise during the next few years, as drought-weakened trees succumbed to bark beetle infestations in their thin-barked upper branches.[107] Because the drought killed an enormous number of the common pine species and firs that surround the discrete groves of giant sequoias, a wildfire of catastrophic intensity, the Castle Fire, swept through the region in 2020. The innate fire resistance of even the tallest and sturdiest sequoias[108] could not protect the groves. A tenth of the entire native population of giant sequoia is estimated to have been killed.[109] An interagency Giant Sequoia Lands Coalition formed in 2021, reporting that "giant sequoias are known for their resistance to insects and disease and their fire-adapted life cycle, however the 2012–2016 drought appears to have been a tipping point for giant sequoias and other Sierra Nevada mixed-conifer forests."[110]

Forest understory plants[]

Seeds of Florida torreya ripen to a purple color at this horticultural planting in eastern North Carolina.

Natural and healthy forests include a diversity of native understory plants. Because understory plants require a healthy canopy of overstory trees, assisted migration of understory plants is only occasionally mentioned in the forestry publications on the topic of climate adaptation.[39]

Assisted migration of forest understory plants[]

Two genera, Taxus and Torreya, within the ancient yew family called Taxaceae have been proposed as important understory components for their value in providing ecosystem services for forest restoration and climate adaptation projects in China. The authors conclude, "Taxaceae species can contribute to generate structurally complex stands of increased value for biodiversity and increased stability, hereby also contributing to climate change mitigation as well as other important ecosystem services. Further, using rare Taxaceae species in reforestation will also help conserve these rare species in a changing future."[111] Both genera have distinct species native to the western and eastern regions of North America. Only Florida torreya is critically endangered, however, and a citizen group called Torreya Guardians has been planting this species hundreds of kilometers north of its native range.[112][113][114]

Inadvertent assisted migration of ornamental understory plants[]

The evergreen southern magnolia is native to the southeastern USA and bears the typical magnolia structure of fruit with fleshy seeds.

Native species of Magnolia in eastern North America have been widely planted for their ornamental beauty in horticultural contexts far north of their native ranges. Although nearly half of all Magnolia species are threatened globally, human-assisted movement and cultivation of some species have led to their survival and expansion far beyond native range. The ongoing naturalization of beyond-range Magnolia taxa has been associated with climate change.[115] Botanists have documented examples where such plantings have fully naturalized — that is, where nearby seedlings and saplings suggest that not only were viable seeds produced, but the habitat and climate were favorable for seedlings to establish and grow. These are examples of inadvertent assisted migration. Botanist Todd Rounsaville suggests, "There is an important need to study and report the ecological processes of tree naturalization in novel environments to guide policy making and forest management relating to climate change, species range-shifts, and assisted migration."[115]

One example is the southern magnolia, Magnolia grandiflora. This small, evergreen tree is widely planted as an ornamental as far north as New England and Michigan. The tree tolerates those regions, which are far north of its native range along the coast of southeastern USA. But no instances of full naturalization have been documented at such high latitudes. In 2011 botanists documented naturalization near Chapel Hill, North Carolina, which is more than a hundred kilometers beyond and inland of its northernmost native range in North Carolina.[116]

Magnolia tripetala in the Morris Arboretum

Bigleaf magnolia, Magnolia macrophylla, another widely planted ornamental, has been documented as fully naturalized in the state of Connecticut. This is 130 kilometers northeast of its historically native range. Because climate models project severe contraction in the southerly portion of its native range, William Moorhead recommends that "additional sites for this conspicuous species should be sought in the area of similar climate conditions between Long Island and central Connecticut, such as along the northern coast of Long Island Sound and in the lower Hudson Valley."[117]

Umbrella magnolia, Magnolia tripetala, has been documented as naturalized in more than a dozen locations well north of its historic native range. Dispersal problems apparently limited its post-glacial range expansion to regions south of the farthest extent of continental ice. While horticultural plantings northward into Massachusetts as early as the 18th century demonstrated species tolerance for the then-climate, only recently have these old horticultural plantings begun to extend offspring into adjacent suitable habitat — becoming quite populous in even full-canopy forests, according to botanists Jesse Bellemare and Claudia Deeg. They write, "The pattern of relatively synchronous escape and establishment of this southern tree species in the last 20 to 30 years seems most consistent with a link to recent climatic warming in the northeastern US."[118]

References[]

  1. ^ a b "Assisted Migration". Natural Resources Canada. Government of Canada. 27 June 2013. Retrieved 20 July 2021.
  2. ^ Beardmore, Tannis; Winder, Richard (January 2011). "Review of science-based assessments of species vulnerability: Contributions to decision-making for assisted migration". The Forestry Chronicle. 87 (6): 745–754. doi:10.5558/tfc2011-091.
  3. ^ Hällfors, Maria H; et al. (July 2014). "Coming to Terms with the Concept of Moving Species Threatened by Climate Change – A Systematic Review of the Terminology and Definitions". PLOS ONE. 9 (7): e102979. Bibcode:2014PLoSO...9j2979H. doi:10.1371/journal.pone.0102979. PMC 4108403. PMID 25055023.
  4. ^ O'Neill, Gregory A; et al. "Assisted Migration to Address Climate Change in British Columbia: Recommendations for Interim Seed Transfer Standards (2008)" (PDF). Ministry of Forests and Range. Government of British Columbia. Retrieved 22 July 2021.
  5. ^ a b "Climate-Based Seed Transfer". Ministry of Forests and Range. Government of British Columbia. Retrieved 20 July 2021.
  6. ^ Leech, Susan March; Almuedo, Pedro Lara; O'Neill, Greg (2011). "Assisted Migration: Adapting Forest Management to a Changing Climate". BC Journal of Ecosystems and Management. 12 (3): 18–34.
  7. ^ a b c Aitken, Sally N; Yeamam, Sam; Holliday, Jason A; Wang, Tongli; Curtis-McLane, Sierra (25 January 2008). "Adaptation, migration or extirpation: Climate change outcomes for tree populations". Evolutionary Applications. 1 (1): 95–111. doi:10.1111/j.1752-4571.2007.00013.x. PMC 3352395. PMID 25567494.
  8. ^ a b Aubin, Isabelle; et al. (January 2011). "Why we disagree about assisted migration: Ethical implications of a key debate regarding the future of Canada's forests". The Forestry Chronicle. 87 (6): 755–765. doi:10.5558/tfc2011-092.
  9. ^ "Northern Forest Affected by Global Warming". NASA Earth Observatory. 22 March 2006. Archived from the original on 2 February 2009.
  10. ^ Prasad, Anantha; Pedlar, John; Peters, Matt; McKenney, Dan; Iverson, Louis; Matthews, Steve; Adams, Bryce (September 2020). "Combining US and Canadian forest inventories to assess habitat suitability and migration potential of 25 tree species under climate change". Diversity and Distributions. 26 (9): 1142–1159. doi:10.1111/ddi.13078. ISSN 1366-9516.
  11. ^ "Annual Regional Temperature Departures". Environment Canada. Archived from the original on January 18, 2012. Retrieved April 2, 2012.
  12. ^ Hogg, E.H.; P.Y. Bernier (2005). "Climate change impacts on drought-prone forests in western Canada". Forestry Chronicle. 81 (5): 675–682. doi:10.5558/tfc81675-5.
  13. ^ Frelich, Lee E; Montgomery, Rebecca A; Reich, Peter B (April 2021). "Seven Ways a Warming Climate Can Kill the Southern Boreal Forest". Forests. 12 (5): 560. doi:10.3390/f12050560.
  14. ^ Jump, A.S.; J. Peñuelas (2005). "Running to stand still: Adaptation and the response of plants to rapid climate change". Ecology Letters. 8 (9): 1010–1020. doi:10.1111/j.1461-0248.2005.00796.x. PMID 34517682.
  15. ^ a b McLane, Sierra C; Aitken, Sally N (2012). "Whitebark pine (Pinus albicaulis) assisted migration potential: testing establishment north of the species range". Ecological Applications. 22 (1): 142–153. doi:10.1890/11-0329.1. PMID 22471080.
  16. ^ St George, Zach (17 June 2021). "As Climate Warms, a Rearrangement of World's Plant Life Looms". Yale Environment 360. Retrieved 25 July 2021.
  17. ^ Allen, Craig D; Breshears, David D; McDowell, Nate G (August 2015). "On underestimation of global vulnerability to tree mortality and forest die-off from hotter drought in the Anthropocene". Ecosphere. 6 (8): art129. doi:10.1890/ES15-00203.1.
  18. ^ Frelich, Lee E; Reich, Peter B (September 2009). "Review: Will environmental changes reinforce the impact of global warming on the prairie– forest border of central North America?". Frontiers in Ecology and the Environment. 8 (7): 371–378. doi:10.1890/080191. hdl:11299/175075.
  19. ^ Seidman, Lila (7 October 2021). "Hundreds of giant sequoias may have burned to death in KNP Complex, Windy fires". Los Angeles Times.
  20. ^ Canon, Gabrielle (3 October 2021). "Giant sequoias and fire have coexisted for centuries. Climate crisis is upping the stakes". The Guardian.
  21. ^ a b Markham, Lauren (November 2021). "Can We Move Our Forests in Time to Save Them?". Mother Jones.
  22. ^ a b Sáenz-Romero, Cuauhtémoc; et al. (August 2016). "Review: Assisted migration of forest populations for adapting trees to climate change". Revista Chapingo Serie Ciencias Forestales y del Ambiente. 22 (3). doi:10.5154/r.rchscfa.2014.10.052.
  23. ^ a b c Williams, Mary I; Dumroese, R Kasten (4 July 2013). "Preparing for Climate Change: Forestry and Assisted Migration". Journal of Forestry. 111 (4): 287–297. doi:10.5849/jof.13-016.
  24. ^ a b c Handler, Stephen; Pike, Carrie; St Clair, Brad; Abbotts, Hannah; Janowiak, Maria. "Assisted Migration: Synthesis prepared for the USDA Forest Service Climate Change Resource Center". Climate Change Resource Center. U.S. Forest Service. Retrieved 21 July 2021.
  25. ^ Aitken, Sally N; Bemmels, Jordan B (July 2015). "Review and Syntheses: Time to get moving: Assisted gene flow of forest trees". Evolutionary Applications. 9 (1): 271–290. doi:10.1111/eva.12293. PMC 4780373. PMID 27087852.
  26. ^ Vose, James M; Peterson, David L; Patel-Weynand, Toral. "Effects of Climatic Variability and Change on Forest Ecosystems: A Comprehensive Science Synthesis for the U.S. Forest Sector (Technical Report, December 2012)" (PDF). U.S. Forest Service. Retrieved 21 August 2021.
  27. ^ a b c Pedlar, John H; et al. (September 2012). "Placing Forestry in the Assisted Migration Debate". BioScience. 62 (9): 835–842. doi:10.1525/bio.2012.62.9.10.
  28. ^ "Forum: Assisted Migration of an Endangered Tree" (PDF). Wild Earth. 14. Winter 2004. Retrieved 20 July 2021.
  29. ^ McLachlan, Jason S; Hellmann, Jessica J; Schwartz, Mark W (26 March 2007). "A Framework for Debate of Assisted Migration in an Era of Climate Change". Conservation Biology. 21 (2): 297–302. doi:10.1111/j.1523-1739.2007.00676.x. PMID 17391179.
  30. ^ Mueller, Jillian M; Hellmann, Jessica J (28 June 2008). "An Assessment of Invasion Risk from Assisted Migration". Conservation Biology. 22 (3): 562–567. doi:10.1111/j.1523-1739.2008.00952.x. PMID 18577085.
  31. ^ Guldberg, O Hoegh; et al. (18 July 2008). "Policy Forum: Assisted Colonization and Rapid Climate Change". Science. 32 (5887): 345–346. doi:10.1126/science.1157897. PMID 18635780. S2CID 206512777.
  32. ^ Seddon, Philip J (31 August 2010). "From Reintroduction to Assisted Colonization: Moving along the Conservation Translocation Spectrum". Restoration Ecology. 18 (6): 796–802. doi:10.1111/j.1526-100X.2010.00724.x.
  33. ^ Barlow, Connie. "Assisted Migration or Assisted Colonization: What's In a Name? Chronological History of the Debate on Terminology". Torreya Guardians. Retrieved 20 July 2021.
  34. ^ Davis, Margaret B (October 1989). "Lags in vegetation response to greenhouse warming" (PDF). Climatic Change. 15 (1–2): 75–82. Bibcode:1989ClCh...15...75D. doi:10.1007/bf00138846. S2CID 154368627.
  35. ^ Davis, Margaret B; Shaw, Ruth B (27 April 2001). "Special Reviews: Range shifts and adaptive responses to Quaternary climate change". Science. 292 (5517): 673–679. doi:10.1126/science.292.5517.673. PMID 11326089.
  36. ^ Petit, Remy J; et al. (August 2004). "Review: Ecology and genetics of tree invasions: from recent introductions to Quaternary migrations". Forest Ecology and Management. 197 (1–3): 113–137. doi:10.1016/j.foreco.2004.05.009.
  37. ^ Seliger, Benjamin J; McGill, Brian J; Svenning, Jens-Christian; Gill, Jacqueline L (November 2020). "Widespread underfilling of the potential ranges of North American trees". Journal of Biogeography. 48 (2): 359–371. doi:10.1111/jbi.14001. S2CID 228929332.
  38. ^ Brodie, Jedediah F; et al. (30 April 2021). "Policy Forum: Global policy for assisted colonization of species". Science. 372 (6541): 456–458. doi:10.1126/science.abg0532. PMID 33926936. S2CID 233448828.
  39. ^ a b c Winder, Richard S; Nelson, Elizabeth A; Beardmore, Tannis (January 2011). "Ecological implications for assisted migration in Canadian forests". The Forestry Chronicle. 87 (6): 731–744. doi:10.5558/tfc2011-090.
  40. ^ Ste-Marie, Catherine; et al. (January 2011). "Assisted migration: Introduction to a multifaceted concept". The Forestry Chronicle. 87 (6): 724–730. doi:10.5558/tfc2011-089.
  41. ^ Nagel, Linda M.; Palik, Brian J.; Battaglia, Michael A.; D'Amato, Anthony W.; Guldin, James M.; Swanston, Christopher W.; Janowiak, Maria K.; Powers, Matthew P.; Joyce, Linda A.; Millar, Constance I.; Peterson, David L. (2017-05-01). "Adaptive Silviculture for Climate Change: A National Experiment in Manager-Scientist Partnerships to Apply an Adaptation Framework". Journal of Forestry. 115 (3): 167–178. doi:10.5849/jof.16-039. ISSN 0022-1201.
  42. ^ Minteer, Ben A; Collins, James P (October 2010). "Communication: Move it or lose it? The ecological ethics of relocating species under climate change". Ecological Applications. 20 (7): 1801–1804. doi:10.1890/10-0318.1. PMID 21049870.
  43. ^ Wilkening, Jennifer L; et al. (1 December 2015). "Alpine biodiversity and assisted migration: The case of the American pika (Ochotona princeps)". Biodiversity. 16 (4): 224–236. doi:10.1080/14888386.2015.1112304. S2CID 131656767.
  44. ^ Gray, Laura K; et al. (2011). "Assisted migration to address climate change: Recommendations for aspen reforestation in western Canada". Ecological Applications. 21 (5): 1591–2103. doi:10.1890/10-1054.1. PMID 21830704.
  45. ^ Swanston, Chris; Janowiak, Maria. "Forest Adaptation Resources: Climate Change Tools and Approaches for Land Managers (2012)" (PDF). Northern Research Station. U.S. Forest Service. Retrieved 5 August 2021.
  46. ^ Schmitt, Kristen M; et al. (May 2021). "Beyond Planning Tools: Experiential Learning in Climate Adaptation Planning and Practices". Climate. 9 (5): 76. doi:10.3390/cli9050076.
  47. ^ Klenk, Nicole L; Larson, Brendon MH (March 2015). "The assisted migration of western larch in British Columbia: A signal of institutional change in forestry in Canada?". Global Environmental Change. 31: 20–27. doi:10.1016/j.gloenvcha.2014.12.002.
  48. ^ a b Buranyi, Stephen. "How British Columbia Is Moving its Trees". Motherboard Tech by Vice, 2016. Vice. Retrieved 21 July 2021.
  49. ^ Brandt, Leslie. "Chicago Wilderness Urban Forest Vulnerability Assessment and Synthesis" (2017). Climate Change Response Framework. U.S. Forest Service. Retrieved 26 July 2021. (Table on page 32.)
  50. ^ "Northern Institute of Applied Climate Science". U.S. Department of Agriculture. U.S. Forest Service. Retrieved 26 July 2021.
  51. ^ Janowiak, Maria; et al. (September 2014). "A Practical Approach for Translating Climate Change Adaptation Principles into Forest Management Actions" (PDF). Journal of Forestry. 112 (5): 424–433. doi:10.5849/jof.13-094.
  52. ^ Bonebrake, Timothy C; et al. (1 June 2017). "Managing consequences of climate-driven species redistribution requires integration of ecology, conservation and social science". Biological Reviews. 93 (1): 284–305. doi:10.1111/brv.12344. PMID 28568902. S2CID 24171783.
  53. ^ Barlow, Connie. "Part 4. Decolonizing Scientific Language". Torreya Guardians. Retrieved 20 July 2021.
  54. ^ Schwartz, Mark W; et al. (12 August 2012). "Managed Relocation: Integrating the Scientific, Regulatory, and Ethical Challenges". BioScience. 62 (8): 732–743. doi:10.1525/bio.2012.62.8.6. S2CID 1623634.
  55. ^ Cooke, Rachel (19 June 2021). "Interview: Robin Wall Kimmerer". The Guardian. Retrieved 21 July 2021.
  56. ^ "Collaborative Projects with Indigenous and Tribal Partners". Center for Native Peoples and the Environment. State University of New York, Syracuse. Retrieved 21 July 2021.
  57. ^ "Tribal Nations Capacity Building". U.S. Climate Resilience Toolkit. U.S. Government. Retrieved 13 September 2021.
  58. ^ a b Tribal Adaptation Menu Team. "Dibaginjigaadeg Anishinaabe Ezhitwaad: A Tribal Climate Adaptation Menu, 2019". Climate Change Response Framework. Northern Institute of Applied Climate Science. Retrieved 21 July 2021.
  59. ^ House, Kelly (3 May 2021). "As northern Michigan warms, scientists bring tree seedlings from the south". Michigan Bridge. Archived from the original on 4 May 2021. Retrieved 21 July 2021.
  60. ^ Zinn, Howard (2005). A people's history of the United States : 1492-present. New York: HarperPerennial. ISBN 0-06-083865-5. OCLC 61265580.
  61. ^ Mucioki, Megan (8 October 2021). "Keeping a Detailed Record of the Changing Climate Could Save this Tribe's Foodways". Civil Eats.
  62. ^ Harrington, Samantha (8 June 2020). "By paying attention, tribes in the Northwoods are leading the way on climate change". Yale Climate Connections.
  63. ^ Risk, Clara; McKenney, Daniel; Pedlar, John; Lu, Pengxin (26 January 2021). "A compilation of North American tree provenance trials and relevant historical climate data for seven species". Scientific Data. 8 (29): 29. Bibcode:2021NatSD...8...29R. doi:10.1038/s41597-021-00820-2. PMC 7838313. PMID 33500421.
  64. ^ McKenney, Dan; Pedlar, John; O'Neill, Greg (March 2009). "Climate change and forest seed zones: Past trends, future prospects and challenges to ponder". The Forestry Chronicle. 85 (2): 258–266. doi:10.5558/tfc85258-2.
  65. ^ Marris, E. (2009). "Forestry: Planting the forest of the future". Nature. 459 (7249): 906–908. doi:10.1038/459906a. PMID 19536238.
  66. ^ "What will climate change do to Alberta Forests?". FRI Research. Retrieved 21 July 2021.
  67. ^ "Tree Improvement and Adaptation Programs" (PDF). Province of Alberta. Retrieved 21 July 2021.
  68. ^ "Managed forests and climate change". Province of Ontario. Retrieved 21 July 2021.
  69. ^ a b Klenk, Nicole L. (2015-03-01). "The development of assisted migration policy in Canada: An analysis of the politics of composing future forests". Land Use Policy. 44: 101–109. doi:10.1016/j.landusepol.2014.12.003. ISSN 0264-8377.
  70. ^ Bennett, Nelson (19 August 2021). "Building climate resilience into our forests". Prince George Citizen (British Columbia).
  71. ^ Dumroese, R Kasten; et al. (August 2015). "Considerations for restoring temperate forests of tomorrow: Forest restoration, assisted migration, and bioengineering". New Forests. 46 (5–6): 947–964. doi:10.1007/s11056-015-9504-6. S2CID 15069899.
  72. ^ "Adaptation". U.S. Department of Agriculture. U.S. Forest Service. 31 July 2017. Retrieved 21 July 2021.
  73. ^ "Plant Species and Climate Profile Predictions". Virginia Technological University. Retrieved 21 July 2021.
  74. ^ Crookston, Nicholas; et al. (August 2010). "Addressing climate change in the forest vegetation simulator to assess impacts on landscape forest dynamics" (PDF). Forest Ecology and Management. 260 (7): 1198–1211. doi:10.1016/j.foreco.2010.07.013.
  75. ^ Prasad, A M; Iverson, L R; Matthews, S; Peters, M. "A Climate Change Atlas for 134 Forest Tree Species of the Eastern United States [database]". Climate Change Tree Atlas. U.S. Forest Service. Retrieved 21 July 2021.
  76. ^ Rehfeldt, Gerald E. (2006). "A spline model of climate for the Western United States". Gen. Tech. Rep. RMRS-GTR-165. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 21 P. 165. doi:10.2737/RMRS-GTR-165.
  77. ^ Iverson, Louis R; Prasad, Anantha M; Matthews, Stephen N; Peters, Matthew (10 February 2008). "Estimating potential habitat for 134 eastern US tree species under six climate scenarios". Forest Ecology and Management. 254 (3): 390–406. doi:10.1016/j.foreco.2007.07.023.
  78. ^ a b Rehfeldt, Gerald E.; Jaquish, Barry C. (March 2010). "Ecological impacts and management strategies for western larch in the face of climate-change". Mitigation and Adaptation Strategies for Global Change. 15 (3): 283–306. doi:10.1007/s11027-010-9217-2. ISSN 1381-2386. S2CID 154285038.
  79. ^ Rehfeldt, Gerald E; Crookston, Nicholas L; Sáenz-Romero, Cuauhtémoc; Campbell, Elizabeth M (1 January 2012). "North American vegetation model for land-use planning in a changing climate: a solution to large classification problems". Ecological Applications. 22 (1): 119–141. doi:10.1890/11-0495.1. PMID 22471079.
  80. ^ Crotteau, Justin S (August 2019). "Initiating Climate Adaptation in a Western Larch Forest" (PDF). Forest Science. 65 (4): 528–536. doi:10.1093/forsci/fxz024.
  81. ^ a b Sáenz-Romero, Cuauhtémoc; O'Neill, Greg; Aitken, Sally N; Lindig-Cisneros, Roberto (2021). "Review: Assisted Migration Field Tests in Canada and Mexico: Lessons, Limitations, and Challenges". Forests. 12 (1). doi:10.3390/f12010009.
  82. ^ Palmer, Clare; Larson, Brendon M H (December 2014). "Should We Move the Whitebark Pine? Assisted Migration, Ethics and Global Environmental Change". Environmental Values. 23 (6): 641–662. doi:10.3197/096327114X13947900181833.
  83. ^ Paz, Fátima (June 18, 2014). "En espera de aprobación de la Profepa por tala ilegal en la Reserva de la Mariposa Monarca". cambiodemichoacan.com.mx
  84. ^ a b Sáenz-Romero, Cuauhtémoc; et al. (20 April 2020). "Review article: Recent evidence of Mexican temperate forest decline and the need for ex situ conservation, assisted migration, and translocation of species ensembles as adaptive management to face projected climatic change impacts in a megadiverse country". Canadian Journal of Forest Research. 50 (9). doi:10.1139/cjfr-2019-0329. S2CID 219081722.
  85. ^ Carbajal-Navarro, Aglean; et al. (6 November 2019). "Ecological Restoration of Abies religiosa Forests Using Nurse Plants and Assisted Migration in the Monarch Butterfly Biosphere Reserve, Mexico". Frontiers in Ecology and Evolution. 7. doi:10.3389/fevo.2019.00421.
  86. ^ Duran, Thelma Gomez (17 January 2022). "Here's how science is trying to conserve the monarch butterfly's forests". Mongabay. Retrieved 25 January 2022.
  87. ^ Anouschka, R; Dymond, Caren C; Mladenoff, David J (November 2017). "Climate change mitigation through adaptation: The effectiveness of forest diversification by novel tree planting regimes". Ecosphere. 8 (11): e01981. doi:10.1002/ecs2.1981.
  88. ^ Muller, Jacob J; Nagel, Linda N; Palik, Brian J (November 2019). "Forest adaptation strategies aimed at climate change: Assessing the performance of future climate-adapted tree species in a northern Minnesota pine ecosystem". Forest Ecology and Management. 451: 117539. doi:10.1016/j.foreco.2019.117539. S2CID 202041429.
  89. ^ "Petawawa Research Forest: Management Goals and Treatments". Adaptive Silviculture for Climate Change. Retrieved 23 July 2021.
  90. ^ Williams, Jeannie (15 January 2020). "Conservation Easement Landowner Newsletter, 2019". Leelanau Conservancy. Retrieved 26 July 2021.
  91. ^ Dennis, Brady (29 April 2020). "In Fast-Warming Minnesota, Scientists Are Trying to Plant the Forests of the Future". The Washington Post. Retrieved 24 July 2021.
  92. ^ "Adaptive Restoration at Stossel Creek". Northwest Natural Resource Group. Retrieved 17 August 2021.
  93. ^ "Stossel Creek Case Study: Adaptive Restoration for Pacific Northwest Forests" (PDF). Northwest Natural Resource Group. Retrieved 17 August 2021.
  94. ^ Braybrook, Rowan (3 August 2021). "Moving Trees: Definitions and Ethics of Assisted Migration". Northwest Natural Resource Group. Retrieved 18 August 2021.
  95. ^ Van der Veken, Sebastiaan; et al. (May 2008). "Garden plants get a head start on climate change". Frontiers in Ecology and the Environment. 6 (4): 212–216. doi:10.1890/070063.
  96. ^ Bellemare, Jesse; Connolly, Bryan; Sax, Dov (June 2017). "Climate Change, Managed Relocation, and the Risk of Intra-Continental Plant Invasions: A Theoretical and Empirical Exploration Relative to the Flora of New England". Rhodora. 119 (978): 73–109. doi:10.3119/16-10.
  97. ^ Marinelli, Janet (19 April 2016). "As World Warms, How Do We Decide When a Plant is Native?". Yale Environment 360.
  98. ^ Lindley, J.Buchholz (1939). "Sequoiadendron giganteum (giant sequoia) description". www.conifers.org. The Gymnosperm Database. Archived from the original on 2011-05-14. Retrieved 2022-01-15.
  99. ^ The redwood forest : history, ecology, and conservation of the coast redwoods. Reed F. Noss, Save the Redwoods League. Washington, D.C.: Island Press. 2000. ISBN 1-55963-725-0. OCLC 41531630.{{cite book}}: CS1 maint: others (link)
  100. ^ "Blue Giant Sequoia". Urban Forest Nursery Inc. Archived from the original on 2021-04-22. Retrieved 2022-01-15.
  101. ^ Barlow, Connie. "Climate, Trees, and Legacy: Videos toward assisted migration". The Great Story. Retrieved 23 July 2021.
  102. ^ Romero, Ezra David (1 September 2020). "Coastal Fog — Or The Lack Of It — Could Be A Wildfire Risk". CapRadio. Retrieved 30 July 2021.
  103. ^ Johnstone, James A; Dawson, Todd E (9 March 2010). "Climatic context and ecological implications of summer fog decline in the coast redwood region". Proc Natl Acad Sci. 107 (10): 4533–4538. Bibcode:2010PNAS..107.4533J. doi:10.1073/pnas.0915062107. PMC 2822705. PMID 20160112.
  104. ^ Fernandez, Miguel; Hamilton, Healy H; Kueppers, Lara M (7 July 2015). "Back to the future: Using historical climate variation to project near-term shifts in habitat suitable for coast redwood". Global Change Biology. 21 (11): 4141–4152. Bibcode:2015GCBio..21.4141F. doi:10.1111/gcb.13027. PMID 26149607.
  105. ^ Francis, Emily J; Asner, Gregory P; Mach, Katharine J; Field, Christopher B (23 June 2020). "Landscape scale variation in the hydrologic niche of California coast redwood". Ecography. 43 (9): 1305–1315. doi:10.1111/ecog.05080. S2CID 225716384.
  106. ^ Dalrymple, Sarah; Winder, Richard; Campbell, Elizabeth M (June 2021). "Guest editorial: Exploring the potential for plant translocations to adapt to a warming world". Journal of Ecology. 109 (6): 2264–2270. doi:10.1111/1365-2745.13715.
  107. ^ Greenfield, Patrick (18 January 2020). "'This is not how sequoias die: It's supposed to stand for another 500 years'". The Guardian.
  108. ^ Quammen, David (2012-12-01). "Forest Giant". nationalgeographic.com. Archived from the original on 3 October 2021. Retrieved 2022-01-14.
  109. ^ Herrera, Jack (3 June 2021). "'Mind-blowing': tenth of world's giant sequoias may have been destroyed by a single fire". The Guardian. Retrieved 23 July 2021.
  110. ^ "Giant Sequoia Lands Coalition". Sequoia and Kings Canyon. U.S. National Park Service. Retrieved 4 August 2021.
  111. ^ Arp Jensen, Ditte; et al. (15 February 2021). "The potential for using rare, native species in reforestation: A case study of yews (Taxaceae) in China". Forest Ecology and Management. 482: 118816. doi:10.1016/j.foreco.2020.118816. S2CID 230577774.
  112. ^ "Editorial: Grows well in sun and warmth — and shade and cold". Nature. 552: 5–6. 4 December 2017. doi:10.1038/d41586-017-07841-1.
  113. ^ Butt, Nathalie; et al. (13 October 2020). "Essay: Importance of species translocations under rapid climate change". Conservation Biology. 35 (3): 775–783. doi:10.1111/cobi.13643. PMID 33047846. S2CID 222320826.
  114. ^ Schipani, Sam (4 October 2018). "Scrappy Group of Citizen Scientists Rallies Around One of World's Rarest Trees". Earth Island Journal.
  115. ^ a b Rounsaville, Todd J (April 2020). "Spatiotemporal recruitment patterns of two introduced Magnolia L. species in a disturbed oak forest". Ecoscience. 27 (3): 165–176. doi:10.1080/11956860.2020.1753311. S2CID 219078456.
  116. ^ Gruhn, Jennifer A; White, Peter S (June 2011). "Magnolia grandiflora L. Range Expansion: A Case Study in a North Carolina Piedmont Forest" (PDF). Southeastern Naturalist. 10 (2): 275–288. doi:10.1656/058.010.0208. S2CID 84283082.
  117. ^ Moorhead, William H (16 March 2018). "Big Leaf Magnolia: A New Addition to the Flora of New England". Rhodora. 119 (980): 349–354. doi:10.3119/17-02. S2CID 90467493.
  118. ^ Bellemare, Jesse; Deeg, Claudia (2015). "Horticultural Escape and Naturalization of Magnolia tripetala in Western Massachusetts: Biogeographic Context and Possible Relationship to Recent Climate Change". Rhodora. 117 (971): 371–383. doi:10.3119/15-04. S2CID 86153619.
Retrieved from ""