Forest and Fire Ecology: Information and Resources

FIRES, INCLUDING HIGH SEVERITY FIRES, ARE GOOD FOR FORESTS

Western conifer forests have evolved with fire and are adapted to it, and in fact need fire. Wildfires, including high-severity fires, create important habitats, lessen the severity and frequency of disease and insect outbreaks, and facilitate seed germination in some plant species. Post-fire forests are one of the most biodiverse and important habitats, and are home to many at-risk, sensitive, and listed species. Post-fire areas are rich in snags (standing dead trees) used for nesting and foraging by many species, including woodpeckers. Fires also stimulate the growth of many species of native wildflowers, certain berries and fungi (such as Morels), and provide habitats for small mammals and forage for deer and elk. Post-fire habitats are among the most important for wildlife, and are unfortunately very rare compared to historic proportions because of continued fire-suppression and post-fire logging.

In studies looking at high-intensity burns, ecologists have found biodiversity equal to, or surpassing, the biodiversity found in old-growth forest. Dr. Richard Hutto, one of the nation’s top ornithologists, found that: “Besides the growing body of evidence that large, infrequent events are ecologically significant and not out of the range of natural variation, an evolutionary perspective also yields some insight into the ‘naturalness’ of severely burned forests… The dramatic positive response of so many plant and animal species to severe fire and the absence of such responses to low-severity fire in conifer forests throughout the U.S. West argue strongly against the idea that severe fire is unnatural. The biological uniqueness associated with severe fires could emerge only from a long evolutionary history between a severe-fire environment and the organisms that have become relatively restricted in distribution to such fires. The retention of those unique qualities associated with severely burned forest should, therefore, be of highest importance in management circles”.

Here is a great summary video about wildfire put together by the group Thriving With Fire and Balance Media:

Born in Fire

 

 

Also check out the video from Wild Lens called “A New Message for Smokey” about the importance of high-intensity wildfires:

More videos and other resources are linked to at the bottom of this page.

PROTECTED FORESTS WITH LITTLE OR NO LOGGING ARE NOT MORE LIKELY TO EXPERIENCE HIGHER SEVERITY WILDFIRE

A recent study from Bradley et al. (2016) challenges USFS assumptions about the fire risk associated with more protected areas—those area that have been less-managed or less-logged, but may still have experienced some degree of fire exclusion (such as Wilderness areas. Streamside riparian areas have also experienced more protection than upland areas for several decades). The authors state:

Mature mixed-conifer forests with natural openings along Coxie Creek in the Camp Lick timber sale. This forest is targeted for logging.

“There is a widespread view among land managers and others that the protected status of many forestlands in the western United States corresponds with higher fire severity levels due to historical restrictions on logging that contribute to greater amounts of biomass and fuel loading in less intensively managed areas, particularly after decades of fire suppression. This view has led to recent proposals—both administrative and legislative—to reduce or eliminate forest protections and increase some forms of logging based on the belief that restrictions on active management have increased fire severity. We investigated the relationship between protected status and fire severity using the Random Forests algorithm applied to 1500 fires affecting 9.5 million hectares between 1984 and 2014 in pine (Pinus ponderosa, Pinus jeffreyi) and mixed-conifer forests of western United States, accounting for key topographic and climate variables. We found forests with higher levels of protection had lower severity values even though they are generally identified as having the highest overall levels of biomass and fuel loading. Our results suggest a need to reconsider current overly simplistic assumptions about the relationship between forest protection and fire severity in fire management and policy”

“Protected forests burn at lower severities: We found no evidence to support the prevailing forest/fire management hypothesis that higher levels of forest protections are associated with more severe fires based on the RF and linear mixed-effects modeling approaches. On the contrary, using over three decades of fire severity data from relatively frequent-fire pine and mixed-conifer forests throughout the western United States, we found support for the opposite conclusion—burn severity tended to be higher in areas with lower levels of protection status (more intense management), after accounting for topographic and climatic conditions in all three model runs. Thus, we rejected the prevailing forest management view that areas with higher protection levels burn most severely during wildfires.”

Furthermore, fire suppression efforts of the past were highly unlikely to have been widespread or effective in remote areas such as those within the project area until recent decades. The 100+ year timeframe the FEA puts forth for fire exclusion is an extreme overestimate. Heyerdahl et al. (2002) note that fire suppression was not effective until recent decades: “…active re suppression by land-management agencies because these efforts were probably not effective until the 1940s–50s when surplus military aircraft became available”.

EFFICACY: LOGGING IS NOT PRACTICAL OR COST EFFECTIVE, AND DOES NOT MAKE COMMUNITIES SAFER

Martha on an old growth Engelmann spruce stump within the Sunrise timber sale area.

The short-term and temporary nature of the perceived fuels reduction benefits from most logging project are not likely to result in meaningful changes to fire intensity, size, or severity. Most projects designed to reduce fuels note that perceived benefits are estimated to affect fire behavior for approximately 20 years. If the estimated effectiveness is only approximately 20 years, then the justification for this project is even more tenuous. For example, Rhodes and Baker (2008) found that: “[u]sing extensive fire records for western US Forest Service lands, we estimate fuel treatments have a mean probability of 2.0-7.9% of encountering moderate-or high-severity fire during an assumed 20-year period of reduced fuels.”

Current forest management is unnecessarily putting firefighters at risk by focusing on remote areas, contrary to peer-reviewed science or common sense. Gibbons et al. (2010) found that defensible space work within 40 meters [about 131 feet] of individual homes effectively protects homes from wildland fire, even intense fire. The authors concluded that the current management practice of thinning broad zones in wildland areas hundreds, or thousands, of meters away from homes is ineffective and diverts resources away from actual home protection, which must be focused immediately adjacent to individual structures in order to protect them. In addition, other studies note that the vast majority of homes burned in wildland fires are burned by slow-moving, low intensity fire, and defensible space within 100-200 feet of individual homes [reducing brush and small trees, and limbing up larger trees, while also reducing the combustibility of the home itself] effectively protects homes from fires, even when they are more intense (Cohen 2000, Cohen and Stratton 2008).

Hutto et al. 2016 note, in relation to climate change, that increased efforts towards fuels reduction would be an untenable emphasis: “Any perceived problem with future changes in fire behavior cannot be solved by redoubling our effort to treat this particular climate change symptom by installing widespread fuel treatments that do nothing to stop the warming trend, and do little to reduce the extent or severity of weather-driven fires (Gedalof et al. 2005). Therefore, fuel management efforts to reduce undesirable effects of wildfires outside the xeric ponderosa pine forest types could be more strategically directed toward creating fire-safe communities….Fuel treatment efforts more distant from human communities may carry the negative ecological consequences we outlined earlier and do little to stop or mitigate the effects of fires that are increasingly weather driven (Rhodes and Baker 2008, Franklin et al. 2014, Moritz et al. 2014, Odion et al. 2014).

In terms of possible future reentry of forests for logging under the guise of ‘fuels reduction’—what is the USFS’s plan for the many thousands of acres of logged areas? Will re-entries be planned across the landscape if wildfire or prescribed fire does not burn the area within 10-35 years? If treatments are only good for 10-15 years, then either re-entry will be needed or logging is a very costly and short-term endeavor that is not effective. It is not feasible to thin the entire landscape every 10-35 years, and if we could accomplish such a misguided feat, it would likely have unacceptable ecological impacts and not be economically viable.

Much peer-reviewed scientific research on mixed-conifer forests has suggested that thinning is likely not needed, effective, nor ecologically beneficial in moist mixed-conifer forest to prevent fire, does not mimic the complex natural fire regime (Noss et al, 2006; Lindenmayer et al. 2009) and threatens to increase fire risk (Lindenmayer et al. 2009). The mixed-conifer forest type is fragile and vulnerable to the chronic negative impacts of industrial commercial logging. Mature and old-growth moist mixed conifer stands have dense, moist interiors and little wind, which inhibit the spread of wildfire (Lindenmayer et al. 2009; Morrison and Smith, 2005; Rhodes, 2007). Large fires are climatically driven and fuels reduction treatments can be insignificant to prevent fire spread under these conditions. Post-fire habitat is significantly degraded by the logging that happens in the name of fuels reduction prior to the fire.

Noss et al. 2006 noted: “One barrier to better use of ecological science is that individuals involved in developing fire policies and practices have tended to be specialists in fire and fuel management, not ecologists, conservation biologists, or other broadly trained scientists. It is not surprising, therefore, that current forest law does not adequately incorporate ecological considerations in its implementation and tends to promote a narrow definition of restoration that focuses almost exclusively on fuels (DellaSala et al. 2004; Schoennagel et al. 2004).”

Downed wood, pools, and hardwoods in natural openings along Bear Creek in the Big Mosquito timber sale.

“True ecological restoration requires the maintenance of ecological processes, native species composition, and forest structure at both stand and landscape scales. Because forests are highly variable over space and time, few universal principles exist for integrating insights from ecology and conservation biology into fire management policies. Nevertheless, one fundamental principle is that managed forests should not only support the desired fire regime but also viable populations of native species in functional networks of habitat (Hessberg et al. 2005). A common-sense conservation goal is to achieve forests that are low maintenance and require minimal repeated treatment. With time, in a landscape of sufficient size, the right end of the restoration continuum (Figure 4) could be reached, where natural fire maintains the system in the desired state. Indeed, wildland fire use is the cheapest and most ecologically appropriate policy for many forests. We envision a future where fire is seen by land managers and the public as the key to healthy forests, but where each forest and each patch of the forest mosaic is recognized for its individuality and managed accordingly. Above all, a guiding principle of forest management should be a precautionary approach that avoids ecological harm.”

LOGGING HARMS MANY AT-RISK AND SENSITIVE WILDLIFE SPECIES

Growing scientific evidence shows that numerous species are at much lower risk from wildfire than previously thought, and some of these species may be more negatively affected by thinning than by wildfire. Example include Olive-sided flycatchers, Canadian lynx, Pacific fisher, Northern spotted owls, flying squirrels, and others (Bond 2015; Bond et al. 2012; Hanson 2015; Moriarty et al. 2016; Manning et al. 2012; Robertson and Hutto 2007). Such research strongly suggests that a greater abundance of caution is needed when considering logging in important wildlife corridors such as riparian areas. For example, Robertson and Hutto (2007) provide evidence for the harmful effects of thinning to some species in their study Is selectively harvested forest an ecological trap for Olive-sided flycatcher? The authors state that: “Human activities that closely mimic the appearance but not the fundamental quality of natural habitats could attract animals to settle whether or not these habitats are suitable for their survival or reproduction. We examined habitat selection behavior and nest success of Olive-sided Flycatchers (Contopus cooperi) in a naturally occurring burned forest and an anthropogenically created habitat type—selectively harvested forest. Olive-sided Fly- catcher density and nestling provisioning rates were greater in the selectively harvested landscape, whereas estimated nest success in selectively harvested forest was roughly half that found in naturally burned forest. Reduced nest success was probably a result of the relatively high abundance of nest predators found in the artificially disturbed forest. These results are consistent with the hypothesis that selectively harvested forest can act as an ‘‘ecological trap’’ by attracting Olive-sided Flycatchers to a relatively poor-quality habitat type. This highlights the importance of considering animal behavior in biodiversity conservation.”

You can watch Dr. Hutto discussing wildfire and birds in the video below. Hutto, R. (posted 2013). Exploring with Dick Hutto. Produced by the Audubon Society.

 

The Forest Service claims that Grand firs and other less fire-resistant trees are present in larger numbers and higher densities across the landscape than they were historically, as a consequence of fire suppression. However, the Forest Service abuses this rationale by applying it overly broadly and aggressively, and uses it as an excuse to extensively log old growth and mature forests– including in ecologically inappropriate areas such as forests with ample evidence of historic mixed-conifer and high-density forests, on north and east facing slopes; deep gulches and narrow valleys; forests on soils that hold more nutrients and moisture (such as ash soils); and other areas that show historic evidence of supporting mixed-conifer forests in general and Grand fir in particular. The Forest Service is using flawed assumptions that lack adequate scientific backing in order to log in ecologically inappropriate and sensitive areas, and to large trees across many thousands of acres—despite the documented deficit in large trees across the landscape and their importance to wildlife.

Large trees are essential for wildlife habitat for numerous species, regardless of the age of trees. The USFS categorizes trees less than 150 years old are “young”, and argues that they do not have to adhere to current Forest Plan prohibitions on logging large trees if they are “young”.  To determine if large trees are less than 150 years old, the USFS uses the Van Pelt guidelines. However, the Van Pelt guidelines are wholly inadequate for  determining the age of Grand fir, as they do not contain guidance that can be used in the field to reliably identify Grand fir older than 150 years. It is large trees that are necessary for wildlife habitat, and those are in extreme deficit across the region due to logging—regardless of age. The Ursus Environmental Assessment on the Deschutes NF discusses the inadequacies of the Van Pelt guidelines for determining age (USFS pg. 77):

“A size or a diameter limit was chosen as the best metric to measure effect on trees that are old or large on the landscape. Other considerations were made, such as using Van Pelt’s guide to identify old grand (white) fir, but due to the characteristics of white/grand fir it was determined to not be an accurate metric. Bark on white/grand fir never develops the thickness of its fire-tolerant associates. The transformation that many trees experience from young gray bark to increasingly more colorful mature bark does not occur with white/grand fir. Even in giant old trees, bark characteristics reveal little about age. Like Douglas fir and western larch, white/grand fir is an opportunist, and has epimoric branch formation. As the stand matures and conditions change around a tree, light penetration may allow new branches to grow where they had been previously lost. Crown condition, tree form, and bark fissures are not an accurate way to tell age. Other than size, there is little else on white/grand fir that indicates age.”

THROUGH AN EVOLUTIONARY LENS

Mature mixed-conifer forest adjacent to Coxie Creek in the Camp Lick sale. The Forest Service is planning both commercial and non-commercial logging in streamside corridors along Coxie Creek.

Mixed-severity fires (including high severity fires) create habitat that is necessary for species within eastern Oregon, suggesting that high severity fires are natural and historically present. High severity fires create important habitat that is very high in biodiversity, and is rare compared to historic norms. The evolutionary history and very existence of native species such as Black-backed woodpeckers suggests that large high-severity fires must regularly occur within their range. We are concerned that current fire suppression efforts through the region will continue to exclude high-severity wildfires (and continue despite project implementation) and threaten the viability of Black-backed woodpeckers and other species that depend on high-severity wildfire.

Hutto et al. (2008) noted that: Without embracing an evolutionary perspective, we run the risk of creating restoration targets that do not mimic evolutionarily meaningful historical conditions, and that bear little resemblance to the conditions needed to maintain populations of native species, as mandated by law (e.g., National Forest Management Act of 1976).” “…[T]he patterns of distribution and abundance for several other bird species (black-backed woodpecker [Picoides arcticus], buff-breasted flycatcher [Empidonax fulvifrons], Lewis’ woodpecker [Melanerpes lewis], northern hawk owl [Surnia ulula], and Kirtland’s warbler [Dendroica kirtlandii]) suggest that severe fire has been an important component of the fire regimes with which they evolved. Patterns of habitat use by the latter species indicate that severe fires are important components not only of higher-elevation and high-latitude conifer forest types, which are known to be dominated by such fires, but also of mid-elevation and even low-elevation conifer forest types that are not normally assumed to have had high-severity fire as an integral part of their natural fire regimes.

In addition, two lines of evidence suggest that it is the more severe fires that are needed to create conditions most suitable for this fire specialist: (1) not only is the black-backed woodpecker restricted to burned forests, but its distribution within burned forests is also relatively restricted to the more severely burned conditions (Kotliar et al. 2002, Smucker et al. 2005, Russell et al. 2007, Hutto 2008); (2) black-backed woodpecker nest sites occur in locations that harbor significantly larger and more numerous trees than occur around randomly selected sites within a burn (Saab and Dudley 1998, Kotliar et al. 2002, Russell et al. 2007). Such nesting locations would be difficult to find in forests maintained as low-density, open, park-like stands due to frequent, low-severity fire.

How could a species evolve to depend on a condition that occurs only infrequently? The answer lies with the distribution and abundance of such fires across space and time. The return interval for severe fire in one location may be several hundred years, but black-backed woodpecker populations persist in a particular re for >6 yr (Saab et al. 2007), so the probability of a severe fire occurring somewhere within an entire watershed in any given 6-year window is very high. Thus, when viewed on a landscape scale, it becomes easy to imagine that a sufficiently mobile plant or animal species (e.g., re morel [Morchella angusticeps]; jewel beetle, Buprestidae; black-backed woodpecker) could become specialized to use a burned forest condition that is ephemeral on a local scale but always present somewhere in the larger landscape.”

FIRE REGIMES AND FOREST STRUCTURE

Multiple convergent lines of evidence suggest that mixed-severity fire, which included large patches of high-severity fire, dominated most western forested landscapes. The extent and influence of fire was much more prevalent than today across the US (USFS National Report on Sustainable Forests, 2010). Evidence that Mixed-severity fires are still within historic range of variability in the west is also provided by the research of Pierce and Meyer (2008). They examined the size and frequency of fire-related debris flows within alluvial fans in Idaho dating back approximately 2,000 years. They found that evidence of small frequent fires historically coincided with multi-decadal climatic cycles of greater moister or drought, and that larger fires corresponded to drought cycles. Today’s high severity fires were found to be within natural range of variability. In addition, fire suppression efforts were highly unlikely to have been widespread or effective in remote areas such as those within the project area until recent decades. The 100+ year timeframe the FEA puts forth for fire exclusion is an extreme overestimate. Heyerdahl et al. (2002) note that fire suppression was not effective until recent decades: “…active re suppression by land-management agencies because these efforts were probably not effective until the 1940s–50s when surplus military aircraft became available”.

Odion et al. (2014) noted, based on extensive literature review of landscape-scale evidence of historical fire severity patterns in Ponderosa pine and mixed-conifer forests:

The USFS is proposing to thin (selectively log) forests within the streamside corridors “RHCAs” in the Ragged Ruby timber sale– including never-logged mixed-conifer forests providing high quality wildlife habitat. We are very concerned that logging will result in the degradation of water quality and wildlife habitat.

“There is widespread concern that fire exclusion has led to an unprecedented threat of uncharacteristically severe fires in ponderosa pine (Pinus ponderosa Dougl. ex. Laws) and mixed-conifer forests of western North America. These extensive montane forests are considered to be adapted to a low/moderate-severity fire regime that maintained stands of relatively old trees. However, there is increasing recognition from landscape-scale assessments that, prior to any significant effects of fire exclusion, fires and forest structure were more variable in these forests. Biota in these forests are also dependent on the resources made available by higher-severity fire.”

“… most forests appear to have been characterized by mixed-severity fire that included ecologically significant amounts of weather-driven, high-severity fire.”

“….paleoecological studies also support mixed-severity fire regimes for the ponderosa pine and mixed-conifer forests. These studies have found charcoal depositions from major fire episodes in ponderosa pine and interior Douglas-fir forests occurring for millennia in the northern Rockies (central Idaho: [100,101]), Klamath [102], Sierra Nevada [103], eastern Oregon Cascades [104], and southwestern USA [105–107]. These major episodes are generally interpreted as large, severe fire events [101– 107].

“The high-severity fire rotations […] do not support the hypothesis that low/moderate-severity fire regimes were predominant in the majority of ponderosa pine and mixed-conifer forests of western North America. In all the large, forest landscapes for which data covering at least 70 years exist, high-severity fire rotations ranged from about 217 to 849 years [57], and were mostly 200–500 years. This is generally less than potential tree lifespans. For combined moderate- and high-severity fires in the eastern Cascades, rotations were 115–128 years”

“The majority of the evidence did not support the low/moderate-severity fire hypothesis, but, instead, supported the alternate hypothesis that mixed-severity fire shaped these forest landscapes. This finding applies to Pacific states ponderosa pine, Jeffrey pine, and California mixed-conifer forests, as well as ponderosa pine and mixed-conifer forests in the eastern Cascades, Rockies and southwestern USA, where low/moderate-severity regimes have often been applied.”

Mature fir and hardwoods along Cougar Creek in the Camp Lick sale.

“In addition, patch sizes of high severity fire in the central Rockies have not increased [58]. Our assessment of high-severity rotations based upon existing literature also revealed a generally lower incidence of high-severity fire in these forests in recent decades…”

“Based on direct observations of fire behavior, high winds (generally 10 m open wind speeds .32–35 kilometers/hr) may subject virtually any conifer forest, regardless of fuel density, to crown fire [108]. Thus, empirical data call into question a major premise of the low/moderate-severity fire regime: that ponderosa pine and mixed-conifer forests may be completely resistant to crown fire. Fire intensity increases with winds, and at winds of .30 km/hr spot fires may be ignited over 1 km ahead of the fire front [109]. The coalescing of separate spot fires with the fire front can further energize wind-driven fire [110,111]. Severe droughts also intensify fires by reducing fuel moisture to extremely low levels, allowing crown fire under less windy conditions [108,112]. Severe drought years throughout much of western North America occurred from 1856 to 1865, 1870 to 1877 and 1890 to 1896 [113]. The extensive high-severity fires of 1910 (the Big Burn in Idaho and Montana), when large areas of drier forests burned at high severity prior to fire exclusion–much of it in ponderosa pine–illustrate how fire behavior that is rare temporally due to extreme climate and weather can dominate in space [1]. Many fire episodes in the charcoal records that exceed modern fires undoubtedly involve combinations of extreme wind, drought, and mass fire.”

“The importance of multiple lines of evidence has been stressed in determining whether mixed-severity fire regimes applied historically [122]. Our results illustrate broad evidence of mixed-severity fire regimes in ponderosa pine and mixed-conifer forests of western North America. Prior to settlement and fire exclusion, these forests historically exhibited much greater structural and successional diversity than implied by the low/moderate-severity model”

“To improve clarity in communication, we propose that ‘‘low/ moderate-severity’’ be applied to those regimes where, as the term implies, high-severity fire is absent. These circumstances appear to be quite rare in the ponderosa pine and mixed-conifer forests of western North America. Therefore, a fire regime with a high-severity component of any amount should not be classified as low/moderate-severity”

“Our findings suggest a need to recognize mixed-severity fire regimes (Table 2) as the predominant fire regime for most of the ponderosa pine and mixed-conifer forests of western North America”

Paula with yews along Bear Creek in the Big Mosquito timber sale.

High intensity wildfires produce unique ecological conditions compared to low intensity fires. High intensity fires are historically natural, and unique, and are integral to the biodiversity of flora and fauna in the region. Favoring lower severity fires across the region in management activities may create unnatural ecological situations that are deleterious to the wildlife, ecological processes, and biodiversity of the area. An example of the importance of post-fire areas occurring after high intensity fires include the findings of Donato et al. (2008):

“[S]evere fires are typically expected to be deleterious to forest flora and development; however, these results indicate that in systems characterized by highly variable natural disturbances (e.g. mixed-severity fire regime), native biota possess functional traits lending resilience to recurrent severe fire. Compound disturbance resulted in a distinct early seral assemblage (i.e. interval- dependent fire effects), thus contributing to the landscape heterogeneity inherent to mixed-severity fire regimes. Process-oriented ecosystem management incorporating variable natural disturbances, including ‘extreme’ events such as SI severe fires, would likely perpetuate a diversity of habitats and successional pathways on the landscape.”

Lady slippers along Bear Creek in the Big Mosquito timber sale.

Historically, mixed-conifer forests were dominated by dense forests (Hessberg et al. 2007). Many of the mixed conifer forests in the west are closed canopy forests that are highly interwoven complex ecosystems that provide high quality habitat for numerous species such as goshawk, marten, Pileated, Lewis woodpecker, and others. Many forests were more dense than commonly recognized by agencies such as the Forest Service.

Williams and Baker have published several studies suggesting that forests in eastern Oregon were considerably more dense than estimates commonly asserted by agencies (Baker 2012; Williams and Baker 2012; and Baker and Williams 2015). There studies have received criticism by some, such as Fule et al. (2014). However, Williams and Baker (2014) provided in-depth responses to these criticisms and clear rationales and defense of their methods; their response was peer-reviewed and published. Their work should not be discounted; rather, it should be considered part of the ongoing conversation and part of the scientific controversy involved in these discussions. Williams and Baker (2014) note:

“Government wildland fire policies and restoration programmes in dry western US forests are based on the hypothesis that high- severity fire was rare in historical fire regimes, modern fire severity is unnaturally high and restoration efforts should focus primarily on thinning forests to eliminate high-severity fire. Using General Land Office (GLO) survey data over large dry-forest landscapes, we showed that the proportion of historical forest affected by high-severity fire was not insignificant, fire severity has not increased as a proportion of total fire area and large areas of dense forest were present historically….. In response, Fulé et al. …suggest that our inferences are unsupported and land management based on our research could be damaging to native ecosystems. Here, we show that the concerns of FE are unfounded. Their criticism comes from misquoting W&B, mistaking W&B’s methods, misusing evidence (e.g. from Aldo Leopold) and missing substantial available evidence. We also update corroboration for the extensive historical high-severity fire shown by W&B. We suggest that restoration programmes are misdirected in seeking to reduce all high-severity fire in dry forests, given findings from spatially extensive GLO data and other sources.”

Mature mixed-conifer forest along Bear Creek in the Big Mosquito timber sale.

The authors respond to each criticism point by point (I encourage people to read it), and then include the following concluding notes:

“The best possible historical baseline for dry forests is likely to come from systematically combining all sources. Past studies that supported the past incomplete historical baseline, which suggested that low-severity fire primarily maintained historical dry forests, were often spatially limited, incomplete samples of larger landscapes. Tree-ring methods can reconstruct to fine scales back to the late 1800s, but are difficult to complete across large landscapes (but see Heyerdahl et al., 2001). Palaeoecological reconstructions can provide key temporal evidence, but are also difficult to replicate across large landscapes. GLO data, in contrast, can be used to develop reconstructions across hundred of thousands to millions of hectares. New findings from GLO data have challenged past findings about the nature of the historical baseline in dry forests, but it is the role of science to continually test past findings. Refining the historical baseline should help avoid misuse of evidence, false narratives, and misdirected restoration and provide a sound scientific foundation for predicting the effects of climatic change on wildfire and forests.”

The conversation and understanding of historic natural conditions continues to evolve. There are still many research gaps; the research that does exist is often from other regions or looks at extremely broad landscapes, consists of very small sample sizes, and/or is contradictory.

We are very concerned that the Forest Service has and will continue to overly broadly apply flawed assumptions regarding their guesses at historical composition of forests,  and apply their target “desired conditions” to special habitats, important wildlife habitats and corridors, and forests that have been substantially less-impacted by the negative effects associated with past management.

Learn more about the beauty and importance of burned forest! Links to great resources:

You can also learn more from the following link:
https://dereklee.scienceblog.com/…/new-documentary-gives-s…/

And you can also listen to a podcast interview discussing the ecological necessity of high-severity fire:
http://wildlensinc.org/film-projects/eyes-on-conservation/…/

Also check out this great presentation on fire ecology and the importance of mixed severity fires by Dominick DelaSalla, PhD.

Here is a wonderful video summarizing the importance of post-fire ecosystems by the Wild Nature Institute:

Other information and links to fire ecology information:

The Chaparral Institute (posted 2014). 11 Reasons Why Burned Forests are Beautiful.
http://californiachaparral.org/wordpress1/2014/08/04/11-reasons-why-burned-forests-are-beautiful/

DellaSala, D. Ecosystem Benefits of Wildfire vs. Post-fire Logging Impacts

Click to access PostFireLoggingFactsheet.pdf

DellaSala, D.; Hanson, C.; Bond, M.; Hutto, R.; Odion, D.; Halsey, R. (2014). Fireside chat: Lessons from Fire Ecology and Post-fire Landscapes. Presentation accessed at: https://prezi.com/embed/efzm54g383n_/ through http://www.geosinstitute.org/banking-on-forests/public-forests/1139-fireside-chat.html

Geos Institute (2013). Open Letter to Members of Congress from 250 Scientists Concerned about Post-fire Logging. http://www.geosinstitute.org/images/stories/pdfs/Publications/Fire/Scientist_Letter_Postfire_2013.pdf

Geos Institute- example pictures of post-fire logging: http://www.geosinstitute.org/images/stories/pdfs/Publications/Fire/SalvageLoggingBiscuitFire.pdf

Hutto, R. 2006. Towards meaningful snag-management guidelines for postfire salvage logging in North American conifer forests. Conservation Biology 20(4) pp 984-993. Accessed online at: http://www.sierraforestlegacy.org/Resources/Conservation/SierraNevadaWildlife/BlackBackedWoodpecker/BBWO-Hutto06.pdf

Sierra Forest Legacy webpage on post-fire logging that contains many good science citations:
http://www.sierraforestlegacy.org/FC_FireForestEcology/FFE_SalvageLoggingScience.php

Wuerthner, G. (2015). Post-fire Logging- a Bad Deal for Forest Ecosystems. Article in Wildlife News. http://www.thewildlifenews.com/2015/03/19/post-fire-logging-a-bad-deal-for-forest-ecosystems/

4-H filming. After the Burn:

 

Other citations for the scientific studies and research on fire regimes, forest density and composition, and the importance of post-fire areas and the negative impacts associated with post-fire logging are listed below. 

Baker, W. L. 2012. Implications of spatially extensive historical data from surveys for restoring dry forests of Oregon’s eastern Cascades. Ecosphere 3(3):23. http://dx.doi.org/10.1890/ES11-00320.1

Baker, W. 2015. Are High-Severity Fires Burning at Much Higher Rates Recently than Historically in Dry-Forest Landscapes of the Western USA? PLoS ONE 10(9): e0136147.

Baker, W. L. and M. A. Williams. 2015. Bet-hedging dry-forest resilience to climate-change threats in the western USA based on historical forest structure. Frontiers in Ecology and Evolution 2:88.

Beschta, R.; Rhodes, J.; Kauffman, B.; Minashall; Karr, J.; Perry, D.; Gresswell, R.; Frissell, C.; Hauer, R.; 2004. Post fire Management on Forested Public Lands of the Western United States. Conservation Biology 18 (4) p. 957–967

Bradley, C.; Hanson, C.; DellaSala, D. 2016. Does increased forest protection correspond to higher fire severity in frequent-fire forests of the western United States? Ecosphere vol 7(10).

Center for Biological Diversity and the John Muir Project (2014). Nourished by Wildfire: The Ecological Benefits of the Rim Wildfire and the Threat of Salvage Logging

Click to access Nourished_by_Wildfire.pdf

Cohen, J.D. 2000. Preventing disaster: home ignitability in the Wildland-Urban Interface. Journal of Forestry 98: 15-21.

Cohen, J.D., and R.D. Stratton. 2008. Home destruction examination: Grass Valley Fire. U.S. Forest Service Technical Paper R5-TP-026b. U.S. Forest Service, Region 5, Vallejo, CA.

D’Antonio, C. 1992. Biological invasions by exotic grasses, the grasses/fire cycle, and global change. Annual Review of Ecological Systems 23 p. 63-87

DellaSala, D. A., Williams, J. E., Williams, C., & Franklin, J. F.; 2004. Beyond Smoke and Mirrors: a Synthesis of Fire Policy and Science. Conservation Biology, 18(4), 976-986

Dillon, G.K., et al. 2011. Both topography and climate affected forest and woodland burn severity in two regions of the western US, 1984 to 2006. Ecosphere 2:Article 130.

Donato, D.C., J.B. Fontaine, W.D. Robinson, J.B. Kauffman, and B.E. Law. 2009. Vegetation response to a short interval between high-severity wildfires in a mixed-evergreen forest. Journal of Ecology 97: 142-154.

Donato, D.; Fontaine, J.; Campbell, J.; Robinson, W.; Kauffman, J.; Law, B.; 2006. Post-wildfire logging hinders regeneration and increases fire risk. Science (New York, N.Y.) 311 (5759) p. 352

Gibbons, P. et al. 2012. Land management practices associated with house loss in wildfires. PLoS ONE 7: e29212.

Hanson, J.; Stuart, J.; 2005. Vegetation responses to natural and salvage logged fire edges in Douglas-fir/hardwood forests. Forest Ecology and Management 214 p. 266–278

Karr, J.; Rhodes, J.; Minshall, W.; Hauer, R.; Beschta, R.; Frissell, D.; 2004. The Effects of Postfire Salvage Logging on Aquatic Ecosystems in the American West. BioScience 54 (11) p. 1029

Keeley, J. 2006. Fire Management Impacts on Invasive Plants in the Western United States. Conservation Biology 20 (2) p. 375-384

Hanson, C.T. , D.C. Odion, D.A. DellaSala, and W.L. Baker. 2009. Overestimation of fire risk in the Northern Spotted Owl Recovery Plan. Conservation Biology 23:1314–1319

Hanson, C.T., D.C. Odion, D.A. DellaSala, and W.L. Baker. 2010. More-comprehensive recovery actions for Northern Spotted Owls in dry forests: Reply to Spies et al. Conservation Biology 24:334–337.

Hanson 2015. John Muir Project of the Earth Island Institute. http://johnmuirproject.org/wp-content/uploads/2014/12/ForestAndFireScienceSynthesisApr2015.pdf

Hanson, C., 2015. Use of higher severity fire areas by female Pacific fishers on the Kern Plateau, Sierra Nevada, California, USA. Wildlife Society Bulletin 39: 497-502.

Haney, A., Apfelbaum, S., and Burris, J., 2008. Thirty years of post-fire succession in a southern boreal forest bird community. The American Midland Naturalist 159: 421-433.

Hessburg P.; Salter, R.; and James, K. 2007. Re-examining fire severity relations in pre- management era mixed conifer forests: inferences from landscape patterns of forest structure. Landscape Ecology 22:5-24.

Heyerdahl, E.; Brubaker, L.; Agee, J.; 2002. Historical re regimes in northwestern USA: annual and decadal climate forcing of historical fire regimes in the interior Pacific Northwest, USA The Holocene 12,5 (2002) pp. 597–604.

Hutto, R., 1995. Composition of Bird Communities Following Stand-Replacement Fires in Northern Rocky Mountain (U.S.A.) Conservation Biology, Vol. 9, No. 5 (Oct., 1995), pp. 1041-1058

Hutto, R. 2008. The ecological importance of severe wildfires: some like it hot. Ecological Applications, 18(8), 2008, pp. 1827–1834

Hutto, R., Conway, C., Saab, V., and J. Walters. 2008. What constitutes a natural fire regime? Insight from the ecology and distribution of coniferous forest birds in North America. Fire Ecology Special Issue, Vol. 4, No. 2.

Hutto, R. L., R. E. Keane, R. L. Sherriff, C. T. Rota, L. A. Eby, and V. A. Saab. 2016. Toward a more ecologically informed view of severe forest fires. Ecosphere 7(2):e01255. 10.1002/ecs2.1255.

Lee, D.E., and M.L. Bond. 2015. Occupancy of California spotted owl sites following a large fire in the Sierra Nevada, California. The Condor, Vol. 117, 2015, pp. 228–236

Lindenmayer, D.; Hunter, M.; Burton, P.; Gibbons, P.; 2009. Effects of logging on fire regimes in moist forests. Conservation letters, doi: 10.1111/j.1755-263X.2009.00080.x. Accessed online at: http://soln.org/wp- content/uploads/2010/01/effects-of-logging-on-fire-regimes-in-moist- forests.pdf

Lindenmayer, D.; Noss, R; 2006. Salvage Logging, Ecosystem Processes, and Biodiversity Conservation. Conservation Biology 20 (4) p. 949-958

Lindenmayer, D.; Burton, P.; Franklin, J.; 2008. Salvage logging and its ecological consequences. Island Press, Washington D.C., USA

Lydersen, J., North, M., and B. Collins, 2014. Severity of an uncharacteristically large wildfire, the Rim Fire, in forests with relatively restored frequent fire regimes. Forest Ecology and Management 328 (2014) 326–334.

Manning, T.; Hagar, J.; McComb, B., 2012. Thinning of young Douglas-fir forests decreases density of northern flying squirrels in the Oregon Cascades. Forest Ecology and Management 264: 115-124.

Moriarty, K.; Epps, C.; Zeilinski, W., 2016. Forest Thinning Changes Movement Patterns and Habitat Use by Pacific Martens. The Journal of Wildlife Management; DOI: 10.1002/jwmg.1060

Morrison, P.H. and H.M. Smith IV. 2005. Fire Regime Condition Classes and Forest Stewardship Planning on the Mt. Hood National Forest. Pacific Biodiversity Institute, Winthrop, WA. 33 p.

Mosgrove, 1980. An Ethnographic History of the Malheur National Forest. USDA.

Noss, R., Franklin, J., Baker, W., Schoennage, T., and P. Moyle, 2006. Managing fire-prone forests in the western United States. Front. Ecol. Environ 2006; 4(9): 481–487

Odion, D.; Hanson, C.; DellaSala, D.; Baker, W.; Bond, M., 2014. Effects of Fire and Commercial Thinning on Future Habitat of the Northern Spotted Owl. The Open Ecology Journal, 2014, 7, 37-51.

Odion D.C., Hanson C.T., Arsenault A., Baker W.L., DellaSala D.A., Hutto R.L., Klenner W., Moritz M.A., Sherriff R.L., Veblen T.T., Williams M.A., 2014. Examining historical and current mixed-severity fire regimes in ponderosa pine and mixed-conifer forests of western North America. PLoS ONE 9: e87852.

Oregon Department of Fish and Wildlife. Accessed 2015. Frequently Asked Questions and Sensitive Species List  http://www.dfw.state.or.us/wildlife/diversity/species/docs/SSL_by_category.pdf

Pacific Northwest Research Station (2017) Science Findings: Woodpecker Woes: the Right Tree Can Be Hard to Find. https://www.fs.fed.us/pnw/sciencef/scifi199.pdf

Pierce, J. and Meyer, G. 2008. Long-term fire history from alluvial fan sediments: the role of drought and climate variability and implications for management of Rocky Mountain Forests. International Journal of Wildland Fire. http://www.publish.csiro.au/wf/wf07027

Pilliod, David S.; Bull, Evelyn L.; Hayes, Jane L.; Wales, Barbara C. 2006. Wildlife and invertebrate response to fuel reduction treatments in dry coniferous forests of the Western United States: a synthesis. Gen. Tech. Rep. RMRS-GTR-173. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 34 p.

Pollock, M., Beechie, T., and Imake, H. (2012). Using reference conditions in ecosystem restoration: an example for riparian conifer forests in the Pacific Northwest. ESA journal, 3(11): 98.

Pollock, M. and Beechie, T. 2014. Does Riparian Forest Thinning Enhance Forest Biodiversity? The Ecological Importance of Downed Wood. Journal of American Waters Resource Association (JAWRA) 50(3): 543-559.

Powers, E.M., J.D. Marshall, J. Zhang, and L. Wei. 2013. Post-fire management regimes affect carbon sequestration and storage in a Sierra Nevada mixed conifer forest. Forest Ecology and Management 291: 268-277.

Raphael, M., Morrison, M., and Yoder-Williams, M., 1987. Breeding bird populations during twenty-five years of postfire succession in the Sierra Nevada. The Condor 89: 614- 626.

Reeves, G. H., Bisson, P. A., Rieman, B. E., & Benda, L. E. (2006). Postfire Logging in Riparian Areas. Conservation Biology, 20(4), 994-1004.

Rhodes, J. J., and W. L. Baker. 2008. Fire probability, fuel treatment effectiveness and ecological tradeoffs in western U.S. public forests. Open Forest Science Journal 1:1–7.

Robertson, B. and Hutto, R., 2007. Is selectively harvested forest an ecological trap for Olive-sided flycatchers? The Cooper Ornithological Society, The Condor 109: 109-121.

Rota, C., 2013. Not all forests are disturbed equally: population dynamics and resource selection of Black-backed Woodpeckers in the Black Hills, South Dakota. Ph.D. Dissertation, University of Missouri-Columbia, MO.

Thompson, J and Spies, T. 2010. Factors associated with crown damage following recurring mixed-severity wildfires and post-fire management in southwestern Oregon. Landscape Ecology 25 pp.775-789

Titus, J.H. ; Householder E.; 2007. Salvage logging and replanting reduce understory cover and richness compared to unsalvaged-unplanted sites at Mount St. Helens, Washington. Western North American Naturalist 67(2) p. 219–231

Wagenbrenner, J.; MacDonald, L.; Coats, R.; Robichaud, P.; Brown, R. (2015). Effects of post-fire salvage logging and a skid trail treatment on ground cover, soils, and sediment production in the interior western United States. Forest Ecology and Management, vol. 335

Williams, M. and W. Baker, 2012. Spatially extensive reconstructions show variable-severity fire and heterogeneous structure in historical western United States dry forests. Global Ecology and Biogeography 750 1.11.

Williams, M. and W. L. Baker, W.; 2014. High-severity fire corroborated in historical dry forests of the western United States: response to Fulé et al. Global Ecology and Biogeography 2014.

Leave a Reply