December 1, 2008

 

Via Federal eRulemaking Portal:

http://www.regulations.gov

 

Public Comments Processing

Attn: FWS-R1-ES-2008-0095

Division of Policy and Directives Management

U.S. Fish and Wildlife Service

4401 N. Fairfax Drive, Suite 222

Arlington, VA 22203

 

Re: Comments on the 90-Day Finding on a Petition to Remove the California, Oregon, and Washington Population of the Marbled Murrelet (Brachyramphus marmoratus) from the List of Endangered and Threatened Wildlife (73 Fed. Reg. 57,314 (October 2, 2008)).

 

Greetings:

 

Please accept the following comments submitted on behalf of The Center for Biological Diversity, Defenders of Wildlife, Oregon Wild, Conservation Northwest, the Environmental Protection Information Center, Klamath-Siskiyou Wildlands Center, Audubon Society of Portland, Kalmiopsis Audubon Society, and Gifford Pinchot Task Force on the U.S. Fish and Wildlife Service's ("FWS" or "Service") 90-day petition finding and initiation of status review for the California, Oregon, and Washington Population of the Marbled Murrelet (Brachyramphus marmoratus) (73 Fed. Reg. 57,314 (October 2, 2008)).

 

In preparing these comments, we have reviewed extensive information related to the ecology and status of the Marbled Murrelet ('the Murrelet'). Based on this review, the Murrelet continues to warrant recognition as a distinct population segment in Washington, Oregon and California because it meets the criteria for discreteness and significance in the Service's policy on recognition of distinct population segments (DPS) (61 Fed. Reg. 4,722 (Feb. 7, 1996)). Because this population continues to face a number of serious threats and has continued to decline since listing, it should be uplisted to endangered. The murrelet is also facing threats in British Columbia and Alaska, where increasing evidence shows severe declines, necessitating listing of the species as threatened in the remainder if its range.

Table of Contents

 

 I. The California, Oregon, and Washington population of the Marbled Murrelet qualifies as a DPS..............3

 

A. The best available scientific information does not support the Service's conclusion that the contiguous U.S. population of the Marbled Murrelet is not discrete..............................3

 

B. The tri-state population of the Marbled Murrelet is discrete under FWS policy......................................5

 

1. Distinct differences exist in regulatory mechanisms across the international border......................5

2. Distinct differences in management of habitat exist across the international border.....................8

3. Distinct differences exist in control of exploitation across the international border......................9

4. There are distinct differences in Conservation Status across the international border.................10

 

C. The tri-state population of the Marbled Murrelet is significant to the taxon to which it belongs.................12

 

D. The Marbled Murrelet should be uplisted to Endangered in the lower 48 states....................................13

 

1. The tri-state population of the Marbled Murrelet has continued to decline since listing............................13

2. Delisting criteria have not been met........................................15

3. Threats to the species are ongoing.........................................16

4. The tri-state DPS is in danger of extinction in light of ESA listing factors...............................................17

 

A. The present or threatened destruction, modification, or curtailment of habitat or range............17

C. Disease or predation...........................................................................................................21

D. Other natural or human caused factors.................................................................................22

1. Oil-Spill Mortality........................................................23

2. Incidental Take in Fisheries..........................................24

3. Reduced and Altered Prey Availability.........................25

4. Disturbance from Motor Craft ....................................27

E. Inadequacy of existing regulatory mechanisms.............................27

 

II. The range wide status of the Marbled Murrelet indicates that the species merits protection under the Endangered Species Act throughout its range.....................31

 

A. The Marbled Murrelet is declining sharply in Alaska and British Columbia..................................32

 

B. The Alaska and British Columbia Marbled Murrelet Population should be listed as threatened under the Endangered Species Act.......................33

 

1. The Alaska and British Columbia population is threatened by four ESA listing factors....................33

 

A. The present or threatened destruction, modification, or curtailment of habitat or range............33

C. Disease or predation................................37

D. Other natural or human caused factors...........................................................39

1. Oil-Spill Mortality...........................................................................39

2. Incidental Take in Fisheries.............................................................41

3. Effects of Sports Fishery on Marbled Murrelets..............................42

4. Reduced and Altered Prey Availability due to Overfishing...............42

5. Reduced and Altered Prey Availability due to Marine Change........43

6. Vessel Disturbance......................45

E. Inadequacy of existing regulatory mechanisms..................46

1. Alaska...................46

2. Canada..................49

 

Literature Cited........................51


I. The California, Oregon, and Washington population of the Marbled Murrelet qualifies as a Distinct Population Segment under Fish and Wildlife Service's Policy for the Recognition of Distinct Vertebrate Population Segments (61 Fed. Reg. 4,722 (Feb. 7, 1996)).

The Policy for the Recognition of Distinct Vertebrate Population Segments states that FWS will consider three elements when determining whether a population segment qualifies as endangered or threatened under the Act: 1. The discreteness of the population segment in relation to the remainder of the species to which it belongs; 2. The significance of the population segment to the species to which it belongs; and 3. The population segment's conservation status in relation to the Act's standards for listing.

A population segment of a vertebrate species may be considered discrete if it satisfies either one of the following conditions:

1. It is markedly separated from other populations of the same taxon as a consequence of physical, physiological, ecological, or behavioral factors. Quantitative measures of genetic or morphological discontinuity may provide evidence of this separation.

2. It is delimited by international governmental boundaries within which differences in control of exploitation, management of habitat, conservation status, or regulatory mechanisms exist that are significant in light of section 4(a)(1)(D) of the Act.

The California, Oregon, and Washington population of the Marbled Murrelet is discrete because of differences between the United States and Canada in regulatory mechanisms, management of habitat, control of exploitation, and conservation status across the international boundary.

A. The best available scientific information does not support the Service's conclusion that the contiguous U.S. population of the Marbled Murrelet is not discrete.

The determination in the 5-year review that the California, Oregon, and Washington Marbled Murrelet population is not discrete is based on an incorrect analysis, is a misinterpretation of DPS policy, contradicts previous findings, and was the result of political interference.

First, the 5-year review was fundamentally flawed. The discreteness analysis in the 5-year review is incorrect because FWS must evaluate differences in regulatory mechanisms as if the Marbled Murrelet were delisted in the United States. The Service acknowledges this in the 90-day finding:

"The Service now believes that the discreteness analysis in the 5-year review was flawed, because it compared current levels of legal protection across the international border, rather than levels of protection that would exist if the marbled murrelet were not listed in the United States. The Service believes that the latter approach is more rational in the context of a 5-year review, because it analyzes discreteness in the same manner as the Service would in an initial listing determination" (FWS 2008, p. 57318, Federal Register: Volume 73, Number 192, October 2, 2008). 

Second, the discreteness analysis in the 5-year review is a misinterpretation of DPS policy. In determining the Murrelet is not discrete, FWS incorrectly interpreted the DPS policy to mean that "differences in control of exploitation, management of habitat, conservation status, or regulatory mechanisms" all had to be "significant in light of section 4(a)(1)(D) of the Act," rather than simply significant differences in any one of these criteria. In particular, FWS concluded that differences in management of habitat are not significant in light of section 4(a)(1)(D), stating: "The differences in management of habitat are not significant in light of section 4(a) (1) (D) of the Act." There are in fact, significant differences in habitat management across the international border, and these are discussed in detail in a following section.

Third, the discreteness analysis in the 5-year review is inconsistent with previous DPS findings. In past interpretations of DPS policy, FWS has determined that populations are discrete across an international boundary for any one of a variety of reasons, including differences in population numbers, regulation, knowledge of the species, and legislation and policy. The Peninsular Bighorn Sheep finding, for example, states:  

"The Service may determine a population to be discrete at an international border where there are significant differences in (1) the control of exploitation; (2) management of habitat; (3) conservation status, or (4) regulatory mechanisms (61 FR 4722). In the case of the Peninsular bighorn sheep, there are significant differences between the United States and Mexico in regard to the species' conservation status" (Federal Register, Vol. 63, No. 52, March 18, 1998, 13134, emphasis added).

 

The Bighorn Sheep and other findings all discuss differences in control, management of habitat, and status without forcing these differences to be significant in light of 4(a)(1)(D).  In the Murrelet 5-year review, FWS inconsistently interpreted the DPS policy to mean that any differences in control, management, or status must be sufficient to actually endanger the species in both countries. This conclusion directly contradicts past findings in which it was specifically known that the species was not threatened or endangered across the border, in particular both the Bighorn Sheep and Pygmy Owl findings, in which it was specifically found that the species were not endangered across the border, as well as the Canada Lynx, Atlantic Salmon, Smalltooth Sawfish and Bull Trout findings, where across the border the species are either not imperiled (lynx) or their status is largely unknown.  

Fourth, FWS Region 1 recognized the Marbled Murrelet as a DPS in the 5-year review, but this decision was reversed due to political interference. In response to an inquiry regarding the involvement of former Deputy Assistant Secretary (DAS) for Fish and Wildlife and Parks, Julie MacDonald, in agency actions or decisions, FWS Region 1 identified actions where "the DAS oversight resulted in an inappropriate change in science that may compromise actions taken in the future, or where the Service's recommendation/decision was changed based on a policy interpretation made by the DAS" (Rabot 2007, p. 1). The reversal in the 5-year review of Region 1's determination that the Marbled Murrelet was in fact a Distinct Population Segment was identified as an instance where a determination based on the best available science was replaced due to political interference, and resulted in an "incorrect" (Rabot 2007, p. 2) analysis: 

"The Service conducted a 5-year review on this species (Brachyramphus marmoratus) in 2004. We used structured decision-making for the 5-year review, working with experts on the marbled murrelet, and making a recommendation that fully considered the risks and uncertainties. The Service recommended that the listing status remain the same, and supported the listing as a distinct population segment (DPS). The Assistant Secretary and the DAS (Deputy Assistant Secretary) subsequently reversed Region 1's DPS determination, stating that the marbled murrelet in the lower 48 was not a DPS, because it did not satisfy the DPS policy's discreteness criteria, and thus was not a listable entity. Since this was based upon an interpretation of the DPS policy made by the Department, the Regional Director signed the 5-year review in support of the Department's position. The Office of the Solicitor review of the proposed delisting rule suggested that the conclusion reached by the Department on the 5-year review was based on an incorrect "discreteness" analysis under the DPS policy, as it compared current levels of legal protection in the United States (Endangered Species Act) and Canada (Species at Risk Act), rather than comparing the levels that would exist if the species were not listed in the United States" (Rabot 2007, p. 2).

Freedom of Information Act documents obtained by Earthjustice reveal that on August 30, 2004, FWS documents stated the tri-state Murrelet population qualified as a DPS, and that on August 31, 2004, Ms. MacDonald circulated a revised document changing the DPS status (Earthjustice 2008).

Thus, FWS's conclusion that the Murrelet is not discrete runs counter to the DPS policy, past DPS designations, is not supported by the best available scientific information based on clear differences in management, regulatory mechanisms, and status of the Murrelet in British Columbia, and resulted from political interference.

B. The California, Oregon, and Washington population of the Marbled Murrelet is discrete under FWS policy.

The contiguous U.S. population of the Marbled Murrelet is discrete because of differences between the United States and Canada in: 1) regulatory mechanisms, 2) management of habitat, 3) control of exploitation, and 4) conservation status.

1. Distinct differences exist in regulatory mechanisms across the international border.

There are distinct differences in existing regulatory mechanisms for the Marbled Murrelet in the United States and Canada. First, SARA provides substantially weaker protection than the ESA, and second, the differences in regulatory mechanisms across the international border would be even greater if the Murrelet were to be delisted in the United States. Further, there are significant differences in regulatory mechanisms for terrestrial habitat management between the United States and Canada.

In terms of existing regulatory mechanisms, there are significant differences between Canada's Species at Risk Act and the ESA. While both laws control the illegal exploitation of Marbled Murrelets from direct take, they differ greatly in terms of Murrelet habitat protection. Compared to the ESA, SARA contains very few provisions for protecting habitat. Under SARA, only the residence (e.g. nest tree) of listed species is protected, and SARA does not provide any habitat protection for the Murrelet beyond the nest tree. SARA only requires that "No person shall damage or destroy the residence" of any threatened or endangered species.  A residence is defined under SARA as: "a dwelling-place, such as a den, nest or other similar area or place, that is occupied or habitually occupied by one or more individuals during all or part of their life cycles, including breeding, rearing, staging, wintering, feeding or hibernation," and has been interpreted to mean only the nest tree when it is occupied. In other words, SARA only protects against the actual harm of a Murrelet and does not provide any protection for its habitat.

 

Further, under the Canadian Species At Risk Act, the residence of listed species is only protected on federal lands. Piatt et al. (2007) state:

 

"In contrast to the protection murrelets have had in the Washington, Oregon, and California area under the ESA, evidence of stand occupancy by murrelets does not prevent forest companies from logging old forests in British Columbia, either on private land or on public leased lands" (p. 130).

 

Moreover, because most Murrelet habitat in British Columbia occurs on provincial land, SARA provides very little protection for Murrelet habitat. In contrast, under the U.S. Endangered Species Act, occupied habitat of threatened and endangered species is protected regardless of land ownership by strict prohibitions against take, which includes habitat protection. Under the ESA, take is defined as "to harass, harm, pursue, hunt, shoot, wound, kill, trap, capture, or collect or attempt to engage in any such conduct." Through regulations, the term "harm" is defined as "an act which actually kills or injures wildlife, and may include significant habitat modification or degradation where it actually kills or injures wildlife by significantly impairing essential behavioral patterns, including breeding, feeding, or sheltering" (USFWS 2008b). For example, ESA protection has typically resulted in the protection of at least the forest stand known or suspected to harbor nesting Murrelets (e.g. 70 acre circles) regardless of land ownership.

 

In addition, the ESA requires the designation of critical habitat for listed species. Under the ESA, critical habitat includes geographic areas that contain the physical or biological features that are essential to the conservation of the species, and federal agencies are required to avoid "destruction" or "adverse modification" of designated critical habitat in federally funded or permitted activities (FWS 2008b). Critical habitat may also include areas that are not occupied by the species at the time of listing but are essential to its conservation. On the contrary, SARA only provides for the identification of critical habitat on non-federal lands as part of the development of a recovery strategy, but the protection of this habitat is entirely discretionary. SARA allows outside groups to petition the Canadian Environment Minister to enact provisions to protect habitat, but the minister is not required to do so, and to date has never enacted additional protections. Although SARA states as a recovery goal that Murrelet habitat loss will be limited to 30%, without critical habitat protection, there is no provision that actually limits the loss of nesting habitat. Furthermore, the details of how SARA will be applied to the Marbled Murrelet have yet to be finalized. Under SARA, the Marbled Murrelet Recovery Strategy is still under review, and Recovery Action Plans are still being developed (Piatt et al. 2007, p. 139).

It is clear that SARA provides substantially less regulatory protection for Murrelet habitat than the ESA. Differences in regulatory mechanisms would be even greater if the Murrelet were to be delisted in the United States. FWS must evaluate the differences in regulatory mechanisms as if the Murrelet were not protected by the ESA, and it is clear that delisting would result in obvious differences in regulatory mechanisms across the international border.

In addition to the differences in regulatory mechanisms concerning Murrelet habitat protection under the ESA and SARA, other regulations concerning forested habitat differ across the international boundary. Canada does not have an overarching forest practices law governing management of its national lands and providing for consideration of wildlife habitat requirements. In contrast, national forests in the United States are governed by the 1976 National Forest Management Act (NFMA) which requires the Forest Service to protect wildlife in national forests and to allow citizens to participate in management decisions. As part of their duty under NFMA to maintain the viability of the Murrelet and other species, the Forest Service created the Northwest Forest Plan (NWFP), which provides substantial protection for Murrelet habitat by creating a network of late-successional reserves. Because of a lack of any legislative mandate to protect habitat for species, similar protections have not been enacted in Canada.  

In Canada, management of forest lands and conservation of wildlife habitat varies depending on Provincial regulations. On Provincial lands in British Columbia, the Murrelet is considered an Identified Wildlife Species under the Forest and Range Practices Act, which calls for habitat protection in Wildlife Habitat Areas (WHAs). WHAs however, provide very little protection for the Murrelet because they are prohibited from impacting the timber base by more than 1%.  This means that no more than 1% of productive mature forest that is suitable for harvest (e.g. doesn't occur on steep slopes, etc.), which comprises the bulk of remaining Murrelet habitat, can be protected.  This 1% cap applies to all identified wildlife species combined, not just the Murrelet, further limiting the availability of protected habitat specifically for the Murrelet. The Forest Practices Board, an independent public watchdog group in British Columbia that investigates how well industry and government meet the intent of British Columbia's forest practices legislation, released a report in 2004 concluding that conservation of Marbled Murrelet habitat under the Forest and Range Practices Act is limited and very slow. While the conservation strategy is still being developed, logging projects are ongoing in Murrelet habitat (Piatt et al. 2007, p. 139).

In previous findings, FWS has recognized populations as discrete based on differences in regulatory mechanisms. For example, the Smalltooth Sawfish finding states: 

"The smalltooth sawfish status review team was unable to identify any mechanisms regulating the exploitation of this species anywhere outside of the U.S. In contrast, several southeastern U.S. states have regulations in place prohibiting fishing for this species. Based on these differences in control of exploitation and regulatory mechanisms, the U.S. population of smalltooth sawfish meets the requirements of discreteness on an international boundary basis." (Federal Register, Vol. 66, No. 73, April 16, 2001, 19414).

Likewise, the Atlantic Salmon finding states:

 "There are substantial differences in the control of exploitation, management of habitat, conservation status, and regulatory mechanisms of Atlantic salmon between the United States and Canada (May, 1993; Baum, 1997). Management and conservation programs in the United States and Canada have similar goals, but differences in legislation and policy support the use of the United States/Canada international boundary as a measure of discreteness for the purposes of evaluating stock status.  Therefore, the Services conclude that the Gulf of Maine population segment of Atlantic salmon satisfies both criteria for demonstrating discreteness, as outlined in the Services' DPS Policy" (Federal Register, Vol. 65, No. 223, November 17, 2000, 69459).

Because of the differences in regulatory mechanisms across the international border, the 3-State Population of the Marbled Murrelet is discrete under DPS policy.

2. Distinct differences in management of habitat exist across the international border.

The United States and Canada differ significantly in management of habitat for the Marbled Murrelet. As detailed above in the discussion on differences in regulatory mechanisms, it is clear that differences in habitat management exist for Murrelet terrestrial habitat between the United States and Canada. Were the Marbled Murrelet to be delisted in the United States, differences would be magnified even further, and as discussed above, FWS must evaluate differences in regulatory mechanisms as if the Marbled Murrelet were delisted in the United States.

 

In the five-year review, FWS argued that there are not significant differences in management between Canada and the United States because similar to listing under the Endangered Species Act, the Murrelet is now listed under Canada's Species at Risk Act ("SARA").  As discussed above, this is not the proper analysis because were the Murrelet to be delisted, this would no longer be the case.  Even if it were a proper analysis, however, the conclusion of the five-year review is not accurate because SARA provides substantially less protection for Murrelet habitat in British Columbia compared to the ESA in the United States. SARA contains very few provisions for protecting habitat. As discussed above, SARA does not provide any habitat protection for the Murrelet beyond the occupied nest tree, and this protection only applies on federal lands, but the majority of Murrelet habitat is on provincial land. SARA does not require the designation of critical habitat and does not contain provisions to actually limit the loss of nesting habitat.

 

In addition to differences in habitat protection under endangered species laws, Murrelet habitat management differs under other policies as well. In the United States, the National Forest Management Act requires that wildlife protection be a priority in forest management. The Northwest Forest Plan, created under the auspices of NFMA, provides late successional reserves for Murrelet protection. Canada, however, does not have an overarching forest practices law. In terms of provincial protection, under British Columbia's Forest and Range Practices Act, habitat protection for all identified wildlife species combined is prohibited from impacting the timber base by more than 1%, severing restricting Murrelet habitat protection.

It is also worth noting that marine habitat protections differ between the United States and Canada, which could affect Murrelet foraging. In the United States, the 1976 Magnuson-Stevens Fishery Conservation and Management Act bans the exploitation of some species of forage fishes, but there is no restriction on the exploitation of forage species in British Columbia (Piatt et al. 2007, p. 93).  

In previous findings, the Service has concluded that differences in habitat management across the international border satisfy the discreteness element of the DPS policy including the Canada Lynx (Federal Register, Vol. 65, No. 58, March 24, 2000, 16052) and Pacific Fisher (Federal Register, Vol. 69, No. 68, April 8, 2004, 18769).

 

It is clear that differences in management exist for Murrelet habitat between the United States and Canada under the ESA and SARA and under other policies. Thus, the Marbled Murrelet satisfies the discreteness criterion of the DPS policy based on differences in habitat management across the international border.

3. Distinct differences exist in control of exploitation across the international border.

Although both the ESA and SARA protect the Marbled Murrelet from direct take, differences in control of exploitation exist across the international border in marine habitat in terms of mortality due to fishing bycatch. Bycatch mortality due to entanglement in fishing nets remains a major threat to the Murrelet in British Columbia (Piatt et al. 2007, p. 90), but has been reduced in the United States. Gill-net fishing has been prohibited or has not occurred for many decades in Oregon and northern California, and was prohibited in Central and Southern California in 2002 (McShane et al. 2004, p. 6-9). Although Murrelet mortality in gillnet fisheries still occurs in Washington, efforts are underway to reduce seabird bycatch, and this is not the case in British Columbia. Piatt et al. (2007) state:

 "There has not been the same intensive research effort made to reduce seabird bycatch in gillnet fisheries in British Columbia as there has been for gillnet fisheries in Washington (Melvin and others 1999)" (p. 92).

In the past FWS has concluded that populations are discrete based on differences in control of exploitation including findings on the Canada Lynx (Federal Register, Vol. 65, No. 58, March 24, 2000, 16052) and Pacific Fisher (Federal Register, Vol. 69, No. 68, April 8, 2004, 18769).

4. There are distinct differences in Conservation Status across the international border.

The Conservation Status of the Marbled Murrelet differs significantly between the United States and Canada. The Marbled Murrelet population in Canada is approximately 2.75 times larger than the Lower 48 population. In addition, threats to the Marbled Murrelet from terrestrial habitat loss and gill-net mortality are higher in British Columbia due to less stringent habitat protections, as discussed above.

In the past, FWS has considered substantial differences in the number of animals across the international border to be sufficient cause for considering the DPS discrete, including the findings for Steller's Eider, Peninsular Bighorn Sheep, and Cactus Ferruginous Pygmy Owl. For example, the Steller's Eider finding states:

"[T]he Alaska breeding population of Steller's eiders is delimited by international boundaries. Within these international boundaries differences in conservation status exist. While available information suggests that the species in Russia also may have declined, population numbers are estimated to range well over 100,000 birds. However, the status of the breeding population in the U.S., as inferred by the contraction of nesting range, is reduced considerably from historic times, despite the existence of regulatory protections and an abundance of seemingly suitable habitat." (Federal Register, Vol. 62, No. 112, June 11, 1997, 31748).

 

FWS acknowledges that "There are differences in population numbers between Canada and Washington, Oregon, and California," but dismisses these differences by saying that "there is no accepted protocol by which these statistics yield a meaningful comparison of conservation status across the border for purposes of the DPS policy."  Specifically, the five year review found that there are roughly 24,000 Murrelets in Washington, Oregon and California compared to roughly 66,000 birds in Canada. 

 

FWS's assertion that there is no accepted protocol is flatly contradicted by past findings and by the fact that there are three times more birds in Canada.  In particular, the finding for the Peninsular Bighorn Sheep concluded:

 

"In the case of the Peninsular bighorn sheep, there are significant differences between the United States and Mexico in regard to the species' conservation status. Information received from the Mexican Government indicates the population in Baja California is not likely to be in danger of extirpation within the foreseeable future because there are significantly more animals there than occur in the United States (Felipe Ramirez, Mexico Institute of Ecology, in litt. 1997). Based on DeForge et al. (1993) there are estimated to be between 780 and 1,170 adult Peninsular bighorn sheep in Baja California, Mexico, north of Bahia San Luis Gonzaga. In addition to the higher population numbers, the Mexican Government has initiated a conservation program for bighorn sheep that should improve the status of these animals. Based on information received from the Mexican Government, components of the conservation program include the involvement of the local people in the establishment of conservation and management units that allow some use of the bighorn sheep while promoting its conservation and recovery. Approximately 1,199,175 ha (485,306 ac) have been included in this program for Peninsular bighorn sheep." (Federal Register, Vol. 63, No. 52, March 18, 1998, 13134)."

 

According to the Bighorn finding, the U.S. had roughly 280 sheep compared to a minimum of 780 in Mexico.  This is very similar to the amount of difference observed between the Murrelet in Canada and the United States, suggesting the observed differences are significant. 

 

Although similar population estimates are not available, FWS relied on greater numbers of animals in findings for both the Cactus Ferruginous Pygmy Owl and Canada Lynx.  In reviewing the determination that the Cactus Ferruginous Pygmy Owl was discrete based on the border, the ninth circuit specifically upheld differences in population numbers as a basis for discreteness, including reference to the Peninsular Bighorn Sheep determination, stating:

 

"The FWS argues that the term "conservation status" means "the number of individuals left in the population." As a consequence, "differences in conservation status" mean "differences in the number of owls" on either side of the border. A court must defer to an agency's interpretation of its own regulations unless it is plainly erroneous. Stinson v. United States, 508 U.S. 36, 45 (1993)… This interpretation is not a post hoc rationalization in the FWS' litigating position, because the FWS has used the term similarly in other listing rules. See, e.g., Endangered Status for the Peninsular Ranges Population Segment of the Desert Bighorn Sheep in So. Calif., 63 Fed. Reg. 13,134, 13,136 (Mar. 18, 1998) (finding "significant differences between the United States and Mexico in regard to the species' conservation status" where "the population in Baja California is not likely to be in danger of extirpation within the foreseeable future because there are significantly more animals there than occur in the United States"). This interpretation also does not reduce the surrounding terms—control of exploitation, management of habitat, regulatory mechanisms—to redundancy or surplusage. See Babbitt v. Sweet Home Chapter of Cmtys. for a Great Or., 515 U.S. 687, 697-98 (1995). We conclude that "conservation status," as used in the discreteness test, is a term of art that lends itself to interpretation by the FWS."

 

Hence, differences in population size alone across the international border qualify the Marbled Murrelet as discrete based on prior interpretations of DPS policy.

C. The California, Oregon, and Washington population of the Marbled Murrelet is significant to the taxon to which it belongs.

The 3-state population of the Marbled Murrelet is significant to the taxon because loss of the discrete population segment would result in a significant gap in the range of the taxon, and because the population differs markedly in its genetic characteristics. 

Loss of the California, Oregon, and Washington discrete population segment would result in a significant gap in the range of the Marbled Murrelet. Loss of the 3-state population would mean the loss of roughly 18% of the linear range of the species, loss of the species from the lower 48 States, and loss of 2.2 million acres of estimated nesting habitat (McShane et al. 2004, p. ES-2).  In the original listing decision, FWS already recognized that the contiguous U.S. population of Marbled Murrelet comprises a significant portion of the species' range, which indicates that it is also significant as defined under the DPS policy.

In the 1992 ruling in Marbled Murrelet v. Lujan, No. C91-522R, slip op. at 4 (W.D. Wash. September 17, 1992) (ER at 905), the district court determined that the tri-state Marbled Murrelet population qualified for listing under the ESA because the Marbled Murrelet habitat in Washington, Oregon, and California constituted a significant portion of the species' range:

[T]he court concludes that, based on the uncontradicted findings that the marbled murrelet qualifies for listing as a threatened species throughout a significant portion of its range within the meaning of the ESA, there is no need to consider the alternative basis of whether the tri-state population is a distinct population segment which might qualify for protection under the ESA (Marbled Murrelet, slip op. at 12 (ER at 913). 

 

The district court also found that all credible science supported finding that the tri-state population was a DPS and that the Service failed "to establish the existence of any scientific dispute on this [DPS] point."  Id. at 12-13 (ER at 913-14).

The California, Oregon and Washington population of Marbled Murrelets also contains marked genetic differences that would be lost should the population be allowed to go extinct.  In particular, Murrelets in central California have been found to contain marked genetic differences and to be of conservation importance (Piatt et al. 2007, p. 42). The central California population is of particular conservation importance because of its location on the periphery of the species range. Piatt et al. (2007) state, "Loss of any of these populations would result in loss of a portion of the species' genetic resources and/or local adaptations, and may compromise its long-term viability" (p. 42).

Because the California, Oregon, and Washington distinct population segment of the Marbled Murrelet represents 18% of the linear range of the species, the entirety of the species' range in the lower 48 states, 2.2 million acres of habitat, and was recognized by FWS as a significant portion of range, its loss would clearly result in a significant gap in the range of the taxon. Further, because the 3-state DPS harbors marked genetic differences, it is clearly significant under DPS policy.

D. The Marbled Murrelet should be uplisted to Endangered in the lower 48 states.

1. The California, Oregon, and Washington population of the Marbled Murrelet has continued to decline since listing. 

Marbled Murrelet populations in the contiguous United States continue to decline and continue to be at risk from the same threats identified at listing. Both monitoring and modeling data indicate that the species is still in decline. Where Marbled Murrelets have been monitored for more than 15 years, data indicate that populations have declined (Huff et al. 2006, p. vii). In Oregon, major population declines (22 to 73 percent) have been documented over a decade or more, with no evidence of increase throughout the breeding range (Huff et al. 2006, p. 9; McShane et al. 2004, p. 6-13). In northern Puget Sound, Washington, Murrelet populations may have declined by 51 percent since the late 1970s (J. Bower) in (Huff et al. 2006, p. 22). In central California, only very small populations of poorly reproducing birds currently occur (McShane et al. 2004, p. ES-2). Populations in Central California appear to be at or near non-viable levels (McShane et al. 2004, p. 6-28).

 

The 5-year review reported that the Murrelet is still declining throughout its listed range: "Demographic modeling with the most recent biological information indicates that the murrelet population is still declining in all 6 Conservation Zones," and that "There is no indication of a population increase in the listed range" (McShane et al. 2004, p. 6-27). Modeling studies of the 3-state DPS suggest population declines of 3 to 7% per year. The 5-year review concludes, "From the available information, long-term survival of the Marbled Murrelet in Washington, Oregon and California is not certain (McShane et al. 2004, p. ES-2).

 

The 3-state population continues to experience low fecundity levels and low breeding success (USFWS 2004, p. 7; Huff et al. 2006, p. 18).  Since the time of listing, both suitable breeding habitat and number of occupied sites have declined throughout the 3-state range, and the decline is expected to continue. McShane et al. (2004) report:

 

"[N]o significant improvements in breeding habitats are expected, such that poor breeding success will likely continue to be major factor affecting populations . . . Breeding populations of murrelets are predicted to continue to decline as areas of old-growth decrease. However, it is likely that murrelet populations will continue to decline even if the amount of nesting habitat remains stable and adult survival unchanged, due to already low levels of breeding success or further reductions in breeding success" (McShane et al. 2004, p. 6-28).

 

McShane et al. (2004) used modeling to evaluate extinction probabilities for Murrelet populations, and calculated an extinction probability of 100% in California, Oregon, and the Western Washington Coast Range during the 21st century (Figure 1). Only the population in Puget Sound has a greater probability of remaining extant than becoming extinct over the next 100 years, with an extinction probability of 25%. Mean population size for the 3-state range was projected to be 45 murrelets after 100 years, with an overall probability of extinction of 16%, and birds surviving only in Puget Sound (McShane et al. 2004, p. 3-58). Populations are likely to become inviable well before reaching extinction. Further, these bleak outcomes are considered to be "optimistic." The authors state:

 

"[T]his calculation may be optimistic because it assumes that the range of population parameters used does not change over 100 years, nesting habitats are not greatly different from today, and mortality from oil spills and gill-nets are similar to recent years. While projected trends and extinction probabilities of zone sub-populations may be alarming to some readers, these predictions are not much different than what could be surmised from existing information and previous modeling efforts (e.g., Beissinger and Nur 1997, Service 1997) . . . [W]e believe that it is doubtful that future projections will differ substantially from those presented here, unless major changes to existing information or substantial new information is uncovered" (p. 3-52).

 

It is clear from both monitoring and modeling data that the contiguous U.S. Marbled Murrelet population continues to decline and warrants uplisting under the ESA from threatened to endangered. Ongoing threats to the Murrelet are discussed below in light of the five listing factors of the Act.

 

 

 

 

 

Figure 1. Population size forecast for the listed range of the marbled murrelet in California, Oregon, and Washington, assuming a 2% annual immigration rate, high-end fecundity, and including gill-net and oil spill mortality. From McShane et al. 2004, Figure 3.5-4, p. 3-51.

2. Delisting criteria for the California, Oregon, and Washington Distinct Population Segment of the Marbled Murrelet have not been met.

The 5-year review concludes that updated information on the species does not indicate that any of the recovery criteria have been met.  The review summary states:

"Does the updated information on the species indicate that any or all of the

recovery criteria for downlisting, delisting or uplisting have been met? No, there is no compelling evidence from the updated information that the trends in estimated population size, densities and productivity have been stable or

increasing in four of the six conservation zones over a 10-year period. It is unclear whether the current management commitments are adequate to protect the murrelet in the six conservation zones for the next 50 years. Attainment of this goal cannot be assured prior to its completion" (USFWS 2004, p. 17-18).

 

The 5-year review clearly states that delisting the 3-state Murrelet population is not warranted:

 

"[D]oes the 5-year review indicate that a change in classification is warranted?

No. The threat situation has not changed such that the murrelet DPS is no longer

likely to become an endangered species within the foreseeable future throughout

all or a significant portion of its range" (USFWS 2004, p. 21).

3. Threats to the species are ongoing and since the time of listing, no threats have been eliminated, some threats have increased, and a new threat has been identified.

None of the threats facing the Marbled Murrelet have been eliminated since the time of listing, and several threats have increased (McShane et al. 2004, p. 6-28).

 

Murrelet nesting habitat continues to be lost, and the effects of past habitat loss continue to negatively affect the Murrelet. Although the rate of habitat loss has slowed since 1992, the amount of old growth forest suitable as breeding habitat and the number of occupied sites have declined throughout the breeding range (McShane et al. 2004, p. 6-29). Further, little habitat is expected to regenerate over the next several decades (Ibid. p 6-30). The 5-year review concludes:

 

 "Overall, threats to marbled murrelet populations in the 3-state area from past and ongoing habitat loss likely remain unchanged since listing due to continued low reproductive success, increased predation, and the low likelihood that additional habitat will develop and be used in the future" (Ibid. p. 6-29 – 6-30).

Oil spills continue to present a serious threat to the Marbled Murrelet. In spite of reductions in oil pollution, oil spills continue to occur and continue to kill relatively large numbers of seabirds, including Murrelets. The 5-year review reports that the rate of oil spill mortality for Murrelets has remained relatively constant (Ibid. 6-31). The risk of catastrophic oil spills is ongoing.

The threat of nest predation on Marbled Murrelets has increased since the time of listing because populations of nest predators have increased. A larger number of species are also now known to be nest predators (Ibid. p. 6-29). Predation on adult Murrelets by raptors may also have increased in some areas, which could have serious demographic consequences because of the influence of adult survivorship on population growth (Ibid. p. 6-29). Threat of extinction in the Santa Cruz Mountains Zone has increased due to very poor breeding success, small population size, increasing predators/predation, and reduced murrelet use of Big Basin Redwoods State Park (Ibid. p. 6-29). The survival of Marbled Murrelets in the Mendocino Zone is dubious. The 5-year review states: "[W]hether or not it (the Zone 5 Mendocino population) can survive for another century is doubtful, especially if adjacent zone populations are reduced to very low numbers" (Ibid. p. 6-29).

 

The potential for increased risk of disease due to the emergence of West Nile Virus and the re-emergence of Newcastle's disease has been identified as a new threat to the Murrelet. The 5-year review states:

 

"The recent emergence of diseases in free-ranging birds in coastal marine systems is an indicator of declining ecological integrity. Diseases in seabirds are expected to increase significantly in the near future as ecological stressors in the marine environment, primarily coastal pollution, increase. Combined with other environmental stressors such as ocean climate changes and habitat loss, diseases may be especially significant with respect to species with declining populations" (Ibid. p. 6-34).

Other threats that have not been alleviated since listing include the risk of unpredictable stochastic events such as wildfires, insect outbreaks, forest diseases, floods, and windstorms; variations in marine prey availability due to climate events; threat from marine contaminants, threat from research efforts; and threat from noise disturbance (Ibid. p. 6-32).

A few threats have been reduced since listing including mortality from gill-net fishing and perhaps reduced loss of occupied sites due to survey error. The 5-Year Review reports that the loss of occupied sites due to survey error has been reduced since listing, but this is debatable at far inland and low-use sites, where the protocol may not be adequate. There has been a decrease in the rate of annual habitat loss, but it is important to note that although the rate of habitat loss has been reduced, habitat loss is ongoing and certainly not improving (Ibid. 6-32). In the listed range since 1992, over 22,000 acres (9,064) hectares of Murrelet habitat has been lost due to timber harvest and natural events, and the Service has consulted on the removal of over 203,000 additional acres suitable Murrelet habitat (McShane et al. 2004, p. 4-64). In Oregon, for example, FWS has allowed logging of known occupied Murrelet sites on Oregon Department of Forestry lands, and more known Murrelet habitat will be lost due to the revised Elliott Forest HCP and the BLM's Western Oregon Plan Revision.

4. The contiguous U.S. population of Marbled Murrelet is in danger of extinction in light of four of the listing factors under the Endangered Species Act.

A. The present or threatened destruction, modification, or curtailment of habitat or

range:

 

The Marbled Murrelet's dependence on old-growth forest is well established. In the forested portion of their breeding range, Murrelet habitat use is strongly correlated with the presence and abundance of mature and old-growth forests, large core areas of old-growth, low amounts of edge and fragmentation, proximity to the marine environment, and increasing forest age and height (McShane et al. 2004, pp. 4-39; Binford et al. 1975, pp. 315-316; Hamer and Nelson 1995, pp. 72-75; Ralph et al. 1995, p. 4, Lank et al. 2003, USFWS 2008, p. 57317). At the stand and tree scale, key components of nesting habitat include large platforms or tree limbs with substrate (moss) and cover, high densities of large trees, layered canopy, and naturally occurring canopy gaps. At the landscape scale, Murrelets are positively associated with large patch size, unfragmented watersheds, and minimal edge (Huff et al. 2006, p. 147). Moreover, at-sea Murrelet abundance is positively associated with inland sites with larger old-growth patches and low levels of fragmentation and isolation regardless of the characteristics of the marine habitat (Burger 2002, Meyer and Miller 2002, Meyer et al. 2002, Miller et al. 2002, Raphael et al. 1995, Raphael et al. 2002b in USFWS 2004, p. 18). At sea, Murrelet abundance is highest where old-growth forests are contiguous with mature second-growth forest (Huff et al. 2006, p. 20).  

 

The past harvest of old-growth forests in the listed range has been the primary contributor to the decline of the Marbled Murrelet (Huff et al. 2006, p. 9, McShane et al. 2004, p. ES-2). For example, in the Pacific Northwest, as much as 90 percent of old-growth forest has been lost, and much of what remains is highly fragmented (Alig et al. 2000, Bolsinger and Waddell 1993, Garman et al. 1999, Hansen et al. 1991, Wimberly and Spies 2000) (in Huff et al. 2006, p. 9).  Habitat loss continues to threaten the Murrelet due to continued logging of habitat, increased predation related to forest fragmentation, and the lack of sufficient time for additional habitat to develop ((McShane et al. 2004, p. 6-30).  The 5-year review concludes:

 

"In most areas within the listed range, murrelets are left with small, isolated stands of older trees for nesting. At present and for the foreseeable future, these remnant populations are struggling to be self-sustaining and may soon become non-viable in Zones 5 and 6 (Central California) and face potential extinction during the next century. It is unrealistic to expect that the species will recover before there is significant improvement in the amount and distribution of suitable nesting habitat" (McShane et al. 2004, p. 6-34).

 

Unfortunately, there has not been improvement in the amount and distribution of suitable nesting habitat. The 5-year review states, "There is no compelling information indicating this situation has improved through the production of significant new suitable nesting habitat since listing" (McShane et al. 2004, p. 18-19).

 

Despite ESA protection, loss of known, occupied Marbled Murrelet nesting habitat is ongoing. The 5-year review concludes:

 

"Overall, the data demonstrate that there has been a continued loss of suitable murrelet habitat due to timber harvest and wildfire since the species was listed in 1992. Further, the time since listing is too short to expect any measurable amount of habitat development. . .  The analysis indicates a continued downward trend in available inland nesting habitat, and most importantly, it shows a loss of known murrelet breeding sites" (McShane et al. 2004, p. 4-107 – 4-108).

 

From 1992 to 2003, over 226,000 acres (91,492 ha) of suitable Murrelet habitat were lost or consulted on for future loss, and an additional 28,119 acres were degraded (McShane et al. 2004, p. 4-107). This loss represents about 10% of estimated suitable habitat, but the proportion of habitat lost is likely greater than 10%, because the amount of estimated suitable nesting habitat is based on using spotted owl habitat as a proxy for Murrelet habitat, but not all spotted owl habitat is suitable for Murrelets (McShane et al. 2004).

Additionally, more than 10,500 potential nest trees were consulted on for removal (McShane et al. 2004, p. 4-107).

 

Of already lost habitat, timber harvest accounted for 24% of the total loss with the majority (75%) of timber harvest occurring in Oregon, and lesser amounts in California and Washington (McShane et al. 2004, p. 4-64). In the Northwest Forest Plan Area, 55,500 acres of habitat were lost to fire and timber harvest, and losses on nonfederal lands were even greater, totaling 139,700 acres. Losses on federal lands were due primarily to fire, whereas losses on nonfederal lands were due primarily to harvest (nearly 100 percent) (Huff et al. 2006, p. 129).

 

Habitat in the 3-state range continues to be lost under Habitat Conservation Plans which involve the potential future removal of thousands of acres of suitable Murrelet habitat and which encompass the largest percent of habitat that could be lost or modified over time. On non-federal lands in Oregon and California, the threat of habitat loss has not changed since listing and remains high (McShane et al. 2004, p. 4-108).

 

Murrelet habitat is also likely to continue to be lost due to stochastic events such as wildfires, wind storms, forest diseases, and insect outbreaks (McShane et al. 2004, p. 6-30). Because of increased forest fragmentation due to timber harvest, the threat of habitat loss due to wind-throw has increased ((McShane et al. 2004, p. 4-79). Global climate change and long-term fire suppression on public lands could lead to more frequent and intense stand-replacing fires, which increases the threat of future Murrelet habitat loss (McShane et al. 2004, p. 4-79). For example, the 2003 Biscuit fire in southern Oregon eliminated approximately 15,000 acres of Murrelet habitat (McShane et al. 2004, p. 6-4). Climate change could also facilitate the outbreak of forest diseases and insect eruptions which represent a threat to murrelet habitat because they can kill complete stands of trees, which both reduces existing suitable habitat and limits the development of future habitat (McShane et al. 2004, p. 4-79).

 

Murrelets are negatively affected by both the amount of lost habitat and the way it is distributed across the landscape because they are highly sensitive to forest fragmentation. Murrelets appear to abandon highly fragmented areas over time (McShane et al. 2004, p. 6-6). The 5-year review states:

 

"[C]hanges in their (Marbled Murrelet) distribution and abundance have occurred in association with habitat loss and forest fragmentation (Service 1997). If murrelets are forced to utilize marginal habitat, nesting success could decline over time, leading to low nesting density and small populations (Raphael et al. 2002b). The fecundity rates of remaining pairs could also decline. Smaller patch size may also affect murrelet nest success and the number of nests, which may ultimately have long-term consequences on population size" (McShane et al. 2004, p. 4-109).

 

Habitat fragmentation can cause population declines greater than those expected from habitat loss alone, due to edge effects which occur when predation rates are higher at habitat edges relative to interiors (Malt and Lank 2007, p. 160). Fragmentation also increases solar radiation and wind in the forest canopy which could affect the distribution of epiphytes, remove moss from nesting platforms, or cause overheating of eggs, chicks, or incubating adults (McShane et al. 2004, p. 6-6).

 

The effects of past habitat loss and fragmentation and ongoing habitat loss and degradation clearly threaten the survival of the Marbled Murrelet. There is scientific consensus that habitat loss is contributing to Murrelet decline. The 5-year review concludes, "The additional habitat loss may result in reduced reproductive success, which is a major factor in the decline of the species" (McShane et al. 2004, p. 4-69). Piatt et al. (2007) state: 

 

"To conclude this section on demographic models, we offer the following perspective on the relevance of models that pay only indirect homage to habitat availability to conservation issues. If a population is declining, not because of relatively subtle changes in birth and death rates of the kind considered here, but rather because some portion of its living space is being modified by humans or nature so as to become wholly unsuitable for reproduction or survival, then the population will decline in direct proportion to the amount of habitat that becomes unusable. In that situation, no mathematical models are really necessary to understand the process" (p. 64).

 

Threats to Marbled Murrelet habitat have increased since the time of listing due to changes in the management of Bureau of Land Management (BLM) lands in Western Oregon and due to ongoing efforts to removal critical habitat for the Murrelet. In July 2008, FWS announced a proposal to remove approximately 254,070 acres of critical habitat in northern California and Oregon. The habitat in Lane and Douglas counties in Oregon slated for loss of protection very likely supports Murrelet populations. In October 2008, the BLM released the Final Environmental Impact Statement for the Western Oregon Plan Revision (WOPR). BLM lands account for 283,000 acres (114,527 ha) or 14% of the suitable Murrelet habitat in the listed range and 38% in Oregon alone (McShane et al. 2004, p. 4-74). The WOPR effectively removes protections for Murrelet habitat under the Northwest Forest Plan (NWFP) and changes the management of 2.6 million acres of land in Western Oregon. Under WOPR, 27 percent of old-growth forests will be harvested within 100 years (BLM 2008). This more than doubles the amount of forest that would have been harvested under the NWFP, from 211 million board feet to 502 million board feet per year, plus another 86 million board feet per year for several decades from habitat improvement projects.

 

Over the course of time, Marbled Murrelet nesting habitat will regenerate, but this process is slow. The 5-year review states, "In general, the earliest possible recovery time for nesting habitat is estimated to be 100-200 years (McShane et al. 2004, p. 4-62). Further, Murrelet use of new habitat may lag for considerable periods. The colonization process is poorly understood and Murrelets display high philopatry which will likely lead to low immigration rates (McShane et al. 2004, p. 4-78).

 

It is clear that the present or threatened destruction, modification, or curtailment of habitat or range places the Marbled Murrelet in danger of extinction, and that the California, Oregon, and Washington Distinct Population Segment should be uplisted to Endangered under the ESA. The 1997 Recovery Plan indicated that the next 50 years would be critical to Murrelet conservation efforts because populations would continue to decline, primarily as a result of loss of nesting habitat (McShane et al. 2004, p. 4-62). It is thus imperative to the long-term survival of the species that ESA protection remain in place. 

 

C. Disease or predation:

 

Murrelet eggs, hatchlings, and adults are all subject to predation. The 5-year review states, "[I]t appears that the threat of predation on murrelets is greater than previously anticipated" (McShane et al. 2004, p. 6-29). Recent studies have shown that the majority of Murrelet nests have failed and that the majority of known nest failures are due to predation (Ibid.). Range-wide studies have revealed nest failure rates due to predation of 68% to 100% in real nests, and 81% to 86% in artificial nests (Hebert and Golightly 2003, Peery et al. in prep., Luginbuhl et al. 2001, Marzluff and Restani 1999 in USFWS 2004, p.19). Similarly, in the 3-state range, most nests (42% - 85%) have failed, and roughly three-quarters of failed nests resulted from predation (based on nests in which it was possible to determine if predation was a factor) (USFWS 2004, p. 10-11).

 

Common ravens (Corvus corax) and Steller's jays (Cyanocitta stelleri) have been implicated as the primary predators of active murrelet nests, and corvids and squirrels (e.g., northern flying squirrel (Glaucomys sabrinus) have been identified as the key predators at artificial nests (Nelson and Hamer 1995, Raphael et al. 2002a in Huff et al. 2006, p. 18). In western North America, populations of several corvid species have increased dramatically as a result of forest fragmentation, increased agriculture, and urbanization (Marzluff et al. 1994 in McShane et al. 2004, p. 2-17).

 

Not only have corvid populations increased, but the number of species known to predate nests has also increased. Known or suspected predators of Murrelet eggs or chicks include sharp-shinned hawks (Accipiter striatus), great horned owls (Bubo virginianus), barred owls (Strix varia), Cooper's hawks (Accipiter cooperi), northwestern crows (Corvus caurinus), American crows (C. brachyrhynchos), and gray jays (Perisoreus Canadensis) (Nelson and Hamer 1995, Nelson 1997, Manley 1999 in McShane et al. 2004, p. 2-16 – 2-17). Artificial nest experiments indicate that predation by squirrels and mice on murrelet eggs and chicks cannot be discounted as a possibility (Luginbuhl et al. 2001, Raphael et al. 2002a, Bradley and Marzluff 2003). The northern flying squirrel (Glaucomys sabrinus), red squirrel (Tamiasciurus hudsonicus), Douglas squirrel (Tamiasciurus douglasi), deer mouse (Peromyscus maniculatus), bushy-tailed woodrat (Neotoma cinerea), and an unidentified mustellid were all recorded attacking plastic eggs and pigeon nestlings (Columba livia) (Marzluff et al. 1999, Flaherty et al. 2000, Luginbuhl et al. 2001, Bradley and Marzluff 2003, in McShane et al. 2004, p. 2-17).

 

Habitat fragmentation increases Murrelet vulnerability to nest predation. Nelson and Hamer (1995) found lower nesting success at sites located closer to forest edges and clearcuts. Burger et al. (2004) found that predation risk at Marbled Murrelet nests was higher near clearcuts and roads than in interior forests, and higher in fragmented landscapes than in relatively intact old-growth forests. Zharikov et al. (2006) found that nest success was lower near regenerating clearcuts, perhaps because of the regeneration of berry-producing shrubs which provide food for Murrelet nest predators. Malt and Lank (2007) found that Murrelet nest disturbances by avian predators were significantly more frequent at hard edges relative to interiors, but less frequent at soft edges, and they did not find any edge effects at natural-edged sites (p. 160). The highest risk of nest predation has been documented in areas close to humans, likely because of food sources for predators such as corvids.   

 

In addition to increased levels of nest predation, the threat of high adult predation by raptors may have increased in some areas due to increased or recovering populations of falcons, eagles, and possibly some hawks. Predation on adult Murrelets likely has more negative consequences for population viability than take of chicks or eggs due to the importance of adult survivorship in demographic models (McShane et al. 2004, p. 6-29). 

 

Not only has the threat posed by predation increased since listing, but so to has the threat posed by the potential spread of avian diseases including West Nile Virus, Newcastle's disease, and others. The 5-year review states:

 

"Murrelets will likely be affected by one or more diseases or biotoxins in the near future because of the cumulative effects of stressors in both their marine and  forest environments. The potential for murrelets to contract West Nile Virus seems high given their nesting habits, the abundance of mosquitoes in forests, the presence of the disease in other forest-nesting bird species, and the proximity of many murrelet nesting areas to open fields, clearcuts, or areas of human activity where corvids are abundant. . . Recent emergence of bacterial, fungal, parasitic, and viral diseases and biotoxins in seabirds poses an increasing threat to Marbled Murrelets" (McShane et al. 2004, p. 3-68 - 3-72).

 

The 5-year review also asserts that the emergence of disease in seabirds could be an indicator of deteriorating ecological conditions:

 

"The recent emergence of diseases in free-ranging birds in coastal marine systems is an indicator of declining ecological integrity. Diseases in seabirds are expected to increase significantly in the near future as ecological stressors in the marine environment, primarily coastal pollution, increase. Combined with other environmental stressors such as ocean climate changes and habitat loss, diseases may be especially significant with respect to species with declining populations" (McShane et al. 2004, p. 6-34).

 

Due to the increased threats posed by predation and disease since the time of listing, the California, Oregon, and Washington Distinct Population Segment of the Marbled Murrelet should be uplisted to Endangered under the ESA.

 

D. Other natural or human caused factors:

 

Other natural or human caused factors which endanger the Marbled Murrelet include oil spills, incidental take in fisheries, reduced and altered prey availability due to a variety of factors including climate change and oceanographic variability, and disturbance from motor craft (McShane et al. 2004, p. 5-12).

 

1. Oil-Spill Mortality

 

Marbled Murrelet mortality from oil pollution is a significant conservation issue in California, Oregon, and Washington (Ohlendorf et al.1978, Burger and Fry 1993, Carter and Kuletz 1995, Service 1997, in McShane et al. 2004, p. 5-16). Since the time of listing, the rate of oil spill mortality of Murrelets has remained relatively constant even though the amount of oil pollution has been reduced. Oil spills continue to occur and kill relatively large numbers of seabirds, including Murrelets. Since listing, the number of oils spills has remained unchanged in Murrelet habitat off the Oregon Coast Range and in Central California, and has increased in habitat off the Siskiyou Coast Range (McShane et al. 2004, p. 5-17). The 5-year review found that the number of Murrelets killed by oil in the listed range has not decreased appreciably since listing, and that because population size has continued to decrease, the proportion of the Murrelet population taken by oil mortality has increased (McShane et al. 2004, p. 5-19).

 

Marbled Murrelets are particularly vulnerable to oil spills because they prefer near-shore habitats and remain at sea most of their lives in near-shore areas near shipping lanes and other sources of oil (Burger 2002 in Piatt et al. 2007, p. 72). Oil spills in California, Oregon, and Washington typically occur close to shore (McShane et al. 2004, p. 5-14).

 

Murrelets can be killed both my major oil spills and by chronic low-level oil pollution. Major oil spills are known to have killed hundreds of Murrelets in Washington (Nestucca and Tenyo Maru) and Oregon (New Carissa) and thousands of Murrelets in Alaska (Exxon Valdez) (McShane et al. 2004, p. 2-16). Although large spills occur infrequently, they have the potential to extirpate Murrelets over entire conservation zones (McShane et al. 2004, p. 5-23). Small spills occur more frequently and have the potential to cause mortality under different wind and sea conditions, particularly when spills occur in foraging areas or during sensitive periods in the breeding season (Piatt et al. 2007, p.75).

 

The 1990 U.S. Oil Pollution Act has generally reduced the number of oil spills, but overall oil tanker and shipping traffic has increased in west coast ports, and seabird mortality due to oil pollution still occurs (McShane et al. 2004, p. 5-38). The 5-year review states, "[I]ncreases in shipping traffic and continued oil pollution will likely outweigh the short-term benefits of increased regulation in the future. Aging sunken vessels also may contribute to additional oil spillage" (McShane et al. 2004, p. 5-23).

Potential off-shore oil development projects also pose a huge threat to Murrelet populations.

 

As Murrelet abundance continues to decline, the effects of oiling mortality of even small numbers of murrelets will increase (McShane et al. 2004, p. 6-31). The chief long-term impact to Murrelet populations from oiling is decreased population size, but decreased breeding success and loss of certain nesting areas also may result if relatively large numbers are killed (Carter et al. 2003 in McShane et al. 2004). Oil impacts are additive to other factors which negatively affect Murrelet populations (Service 1997 in McShane et al. 2004, p. 5-15).

 

2. Incidental Take in Fisheries

 

The incidental capture of birds in gillnets is a well-documented effect of fisheries on Marbled Murrelets (Piatt et al. 2007, p. 78). In the past, gill-net mortality has had significant impacts on some Murrelet populations (Piatt et al. 2007, p. 78). In Oregon and Northern California, gill-net fishing has been prohibited or has not occurred recently, and gill-netting was prohibited in Central California in 2002 (McShane et al. 2004, p. 6-9). In Washington mortality from gill-net fishing is ongoing (McShane et al. 2004, p. ES-2).

 

Marbled Murrelet mortality from coastal gill-net fishing has been considered a significant conservation issue in Puget Sound and the Western Washington Coast Range (McShane et al. 2004, p. 5-23). The best available estimate of annual Murrelet loss from gill-net fishing in Washington is up to 120 birds per year from 1980 to 1992, and 30 birds per year from 1993 to 2003. Gill-net mortality is thought to have been reduced in Washington for several reasons: The Washington Department of Fish and Wildlife has voluntarily implemented a number of practices to limit by-catch from gill-nets in non-tribal fisheries; fishing effort has decreased due to declining salmon runs; alternative fishing gear has been implemented, though whether the alternative gear reduced Murrelet mortality has not been determined (Piatt et al. 2007, p. 88). Despite these measures, some level of gill-net murrelet mortality continues (McShane et al. 2004, p. 5-39). The 5-year review concludes that gill-net mortality in the Puget Sound population has probably "significantly contributed to population declines prior to and after listing" (McShane et al. 2004, p. 5-32). The review cautions that additional closures may be necessary, and that mortality could increase with fishing effort:  

 

"In Zones 1 and 2 (Puget Sound and Western Washington Coast Range) impacts continue and are additive to other problems. We suggest that additional work is needed to better determine current bycatch levels and estimates of past mortality. Additional closures may be needed in nearshore fishing areas where relatively large numbers of murrelets are aggregated and mortality may be highest. While fishing effort is currently low, there is some indication that effort may increase in the near future if salmon stocks rebound" (McShane et al. 2004, p. 5-35).

 

Reduced population size is the chief long-term impact to Murrelet populations from gill-net mortality. Reduced breeding success and loss of use of certain nesting locations could also occur. If only small numbers of birds nest in particular locations, then the loss of a few individuals could lead to loss of use of those nesting areas. McShane et al. (2004) state, "Regardless of the level of impacts, they are additive to other factors negatively affecting murrelet populations" (McShane et al. 2004, p. 5-24).

 

In addition to gill-net bycatch, Marbled Murrelets also can be killed by entanglement with fishing lines or hooking with fishing lures (Carter et al. 1995) (McShane et al. 2004, p. 5-23).

3. Reduced and Altered Prey Availability

 

Reduced and altered prey availability is an ongoing threat to the 3-state Marbled Murrelet population. Many fish populations in Washington, Oregon, and California have been depleted due to overfishing from coastal fisheries (McShane et al. 2004, p. 5-12 – 5-13). In California, the diet of the Marbled Murrelet has shifted over the last century to lower trophic level prey items (such as krill, rockfish, and sand lance) in response to a decrease in the abundance of higher trophic level prey species associated with declines in fisheries (Becker and Beissinger 2005). Murrelets overall may not be significantly affected by current levels of overfishing, but in certain areas local impacts may occur which could affect at-sea Murrelet distribution (McShane et al. 2004, p. 5-12 – 5-13). Reductions and alterations in prey availability have also occurred due to reduced quantity and quality of spawning habitat, pollution of coastal waters, and oceanographic variation due to natural cycles and climate change (Ibid). Changes in marine community composition, whether due to over-fishing or natural variation in the marine environment and patterns of upwelling, have important affects on the prey base and demographic response of Marbled Murrelets" (Piatt et al. 2007, p. 71).

 

Prey availability has direct effects on Marbled Murrelet distribution and productivity (Peery et al. 2004b, Becker and Beissinger 2005 in Piatt et al. 2007, p. 71). Both the abundance and quality of available food is a critical determinant of reproductive success for piscivorous seabirds. Forage fishes differ substantially in fat content, energy density, and quality (Anthony et al. 2000). There can be also be differences in energy density within the same fish species, depending on season, environmental conditions, sex, and/or physical condition (Robards et al 1999a, Anthony et al. 2000 in Piatt et al. 2007, p. 72). The nutritional quality of prey affects rates of delivery to chicks, chick growth, and chick survival (Golet et al. 2000, Suryan et al. 2002, Wanless et al. 2005 in Piatt et al. 2007, p. 72). For seabird chick growth, prey quality may be even more important than prey quantity (Kitaysky et al. 1999; Wanless et al. 2005 in Piatt et al. 2007, p. 72). Prey quality and quantity also affects parental condition and predation rates. If prey quality or quantity is low, parents must spend more time capturing and delivering prey, which is both energetically expensive and increases exposure to raptors and nest predators (Peery et al. 2004b, Piatt et al. 2007, p. 72).

 

Large-scale shifts in marine climate contribute to biological regime shifts which affect the abundance and type of forage fish available for Marbled Murrelets and other seabirds (Anderson and Piatt 1999, Benson and Trites 2002, Chavez et al. 2003 in Piatt et al. 2007, p. 71). Most of the fish species upon which Marbled Murrelets prey consume macro-zooplankton including copepods and euphausiids. Murrelets are thus linked to changes in ocean and climate conditions which influence zooplankton abundance, distribution, and species composition (Zebdi and Collie 1995, Williams and Quinn II 2000, Mackas et al. 1998, 2001, Boldt et al. 2005 in Piatt et al. 2007, p. 72). The El Niño-Southern Oscillation, Pacific Decadal Oscillation, and other factors that affect ocean temperatures and current flows impact prey distribution and alter Murrelet foraging locations and success rates (Ainley et al.1995 in Huff et al. 2006, p. 20).

 

The El Niño-Southern Oscillation (El Niño) occurs when internal ocean waves generated in the tropics infuse warm water along the western North American coast and create corresponding changes in the mixed layer depth. It occurs every four to seven years and lasts for a few to several months. El Niño negatively affects ocean productivity and the food web. Hallowed et al. (2001) found that El Niño affects fish production from California to Washington. El Niño events have been shown to have dramatic effects on seabird productivity as a result of reduced prey availability (Ainley and Boekelheide 1990, Ainley et al. 1995b in McShane et al. 2004, p. 5-9). In strong El Niño events, seabirds forgo breeding entirely, and in weaker ones seabird reproduction is significantly reduced because fewer birds breed, and greater numbers of breeding attempts fail (Ainley et al. 1995b in McShane et al. 2004, p. 5-4 – 5-5).

 

Several studies have shown that El Niño influences the behavior of Marbled Murrelets.

Becker (2001) found changes in the at-sea distribution of murrelets and a shift to higher trophic-level feeding during spring in El Niño years. Burger (2000) found differences in the number of Murrelets detected with radar entering drainages in British Columbia in El Niño versus non-El Niño years, and proposed that this indicated a decrease in breeding activity due to reduced prey availability (McShane et al. 2004, p. 5-13). In warm-water years, Murrelets occur at lower densities in near-shore waters and breeding attempts appear to be significantly reduced (Huff et al. 2006, p. 20).

 

Decadal-scale variation in oceanographic climate also may affect prey availability for Marbled Murrelets. The Pacific Decadal Oscillation (PDO) occurs every 15 to 20 years and changes the depth of the mixed layer. In California, Oregon, and Washington zooplankton and fish populations are negatively affected in the warm phases of the PDO when the mixed layer is deep (Batchelder et al. 2002, Batchelder and Powell 2002, Hallowed et al. 2001). Tri-state seabird populations have also done very poorly during the most recent warm phase of the PDO due to variations in prey availability (Veit et al.1997, Ainley and Divoky 2001 in McShane et al. 2004, p. 5-14).

 

Murrelets in Oregon and Washington are potentially negatively affected by the recent annual occurrence of a "dead zone" in near shore waters, an area where dissolved oxygen levels are so low as to be lethal. A dead zone has been appearing along the Oregon coast each summer since 2002, in response to the anomalous flow of subarctic water into the California Current System. The 2006 dead zone blanketed approximately 70 miles (113 kilometers) of the central Oregon coast and possibly extended another 170 miles (274 kilometers) into Washington waters. This "unprecedented" development has resulted in "mass die-offs of fish and invertebrates within the California Current System" (Grantham et al. 2004, p. 749). Grantham et al. (2004) state, "Our findings highlight the sensitivity of inner-shelf ecosystems to variation in ocean conditions, and the potential impacts of climate change on marine communities" (Grantham et al. 2004, p. 749).

 

Although El Niño and the PDO are natural and cyclic, some climatic changes may be linked to anthropogenic causes (McShane et al. 2004, p. 5-14). Further, fluctuations in oceanic climate may be increasing in frequency (King 2005, Piatt et al. 2007, p. 71). The increased frequency of such events, and their resultant effects on prey availability, are a further stressor for already declining Marbled Murrelet populations.

 

  4. Disturbance from Motor Craft

 

Marbled Murrelets can be negatively affected by traffic in both marine and terrestrial habitat. In the marine environment, several studies have shown that Brachyramphus murrelets are displaced by vessel traffic, and possibly impacted energetically.

Hamer and Thompson (1997) found that nearly 70% of moved a short distance away when a vessel approached (in Piatt et al. 2007, p. 96-97). Kuletz et al. (2003) found that Kittlitz's Murrelets were temporarily displaced by boat traffic and that when boats were present fewer birds made foraging dives and more birds flew off the water. Hentze (2006) found that single birds were more likely to flush, and because single birds are more likely to be breeding individuals, suggested that boat disturbance disproportionately affects breeding birds (in Piatt et al. 2007, p. 96-97). Speckman et al. (2004) found that the approach of small boats caused Marbled Murrelets holding fish (presumably meant for their chicks) to dive and then swallow the fish. This behavioral response could be energetically costly for chick-rearing murrelets and could have implications for reproductive success. Vessel disturbance can increase energy demands on Marbled Murrelets and could lead to reductions in fitness under adverse environmental conditions (Agness 2005 in Piatt et al. 2007, p. 142).

 

In the terrestrial environment, Nelson (1997) reported at least five documented instances of Marbled Murrelet mortality resulting from vehicular collisions. Where nests are located in the vicinity of roads, nesting adults are particularly susceptible to vehicular traffic risk because they typically approach nests from below (Nelson 1997 in McShane et al. 2004, p. 2-19).

 

In sum, the California, Oregon, and Washington Distinct Population Segment of Marbled Murrelet should be uplisted to Endangered under the ESA because of other natural or human caused factors including oil spills, incidental take in fisheries, reduced and altered prey availability due to overfishing, climate change and oceanographic variability, and disturbance from motor craft (McShane et al. 2004, p. 5-12). Each of these factors compounds existing stressors on already-declining populations.

 

E. Inadequacy of existing regulatory mechanisms:

           

The California, Oregon, and Washington Distinct Population Segment of the Marbled Murrelet is currently protected under the Endangered Species Act. Were the DPS to be delisted, however, existing regulatory mechanisms would be completely inadequate to protect the 3-state population on federal, state, or private lands.

 

Federal lands account for the majority of suitable Marbled Murrelet habitat in Washington, Oregon, and California-- approximately 2 million of the 2.2 million estimated total acres of suitable habitat are located on Federal lands (93%) (McShane et al. 2004, p. 4-10). Even with ESA protection, Murrelet habitat loss on federal lands is ongoing. Were the Murrelet to be delisted, this habitat loss would be even greater. Since the time of listing, 55,500 acres of high-quality Murrelet habitat have been lost on federal lands due to timber harvest and wildfire (Huff et al. 2006, p. 126).

 

In the Pacific Northwest, the Marbled Murrelet is afforded some habitat protection on federal lands by the Northwest Forest Plan (NWFP), but without Endangered Species Act protection, the NWFP alone would not adequately protect the Marbled Murrelet. The NWFP is not a species recovery plan and cannot be relied upon to ensure persistence of the Murrelet in the absence of federal level protection. The NWFP created Late-Successional Reserves (LSRs) which theoretically provide breeding habitat for Marbled Murrelets. Not all LSR habitat, however, is actually high-quality Murrelet habitat. Of the 2.8 million habitat-capable forest acres that occur within LSRs, only 36 percent is classified as higher suitability Murrelet habitat (Huff et al. 2006, p. 129). About 8% of Marbled Murrelet habitat on federal lands is not in reserved status and may be logged, and there are significant coastal habitat areas that are not covered by the NWFP.

Loss of Murrelet habitat is ongoing under the NWFP. Section 7 consultations throughout the 3-state area anticipated the removal of 10,537 potential nest trees and degradation of another 50 trees. All but 12 of these trees were removed or degraded in Oregon and Washington, and all but 28 were removed or degraded on USFS and BLM lands covered by the NWFP (McShane et al. 2004, p. 4-69). Further, during the first ten years of the Northwest Forest Plan, low Marbled Murrelet nest success rates prevailed throughout the Plan area (Huff et al. 2006, p. 18). An evaluation of the success of the plan during its first ten years states, "Our baseline estimates indicate that in general the Plan area has low amounts of forest with predicted high odds ratios of providing suitable habitat for Marbled Murrelet nesting" (Huff et al. 2006, p. 91). The protection of Marbled Murrelet nesting habitat on areas outside of the NWFP area is thus extremely important.

 

The BLM's Western Oregon Plan Revision (WOPR) poses a new threat to Murrelet habitat in Western Oregon by removing BLM lands from the management guidance of the NWFP. The WOPR removes LSR designations in Murrelet habitat in the Oregon Coast Range and in the Siskiyou Coast Range. Nearly one-quarter of land in the Oregon Coast Range is managed by Federal agencies, and the majority of high quality suitable murrelet habitat in the Oregon Coast Range occurs on U.S. Forest Service and BLM lands (McShane et al. 2004, p. 4-13). BLM lands account for 283,000 acres (114,527 ha) or 14% of the suitable Murrelet habitat in the listed range and 38% in Oregon alone (McShane et al. 2004, p. 4-74). BLM released the Final Environmental Impact Statement for the Western Oregon Plan Revision in October 2008. The WOPR changes the management of 2.6 million acres of land in Western Oregon, and calls for the harvest of 27 percent of old-growth forests within 100 years (BLM 2008). This obviously poses a major threat to Marbled Murrelet nesting habitat, and more than doubles the amount of forest that would have been harvested under the NWFP, from 211 million board feet to 502 million board feet per year, plus another 86 million board feet per year for several decades from habitat improvement projects. Concerning the instatement of WOPR and the removal of BLM lands from the NWFP, Suzuki and Olson (2008) state, "Hence the role and capacity of these federal lands for biodiversity conservation would be significantly altered" (Suzuki and Olson 2008, p. 1022).

 

Habitat protection on federal lands is of great importance to Marbled Murrelet conservation. Were the 3-state DPS to be delisted, there would be very little protection for the species on private lands. Suzuki and Olson (2008) conclude: "It is likely that any removal of conservation measures on federal lands due to a policy change would not be compensated by the current level of conservation efforts on nonfederal lands" (p. 1035).

 

In the 3-state area, non-federal public lands (State and County) and private ownership account for approximately 8% of the total area of suitable Murrelet habitat; Tribal lands account for less than 1% of the total area of suitable habitat (McShane et al. 2004, p. 4-10). Since the time of listing, Marbled Murrelet habitat has continued to be lost on nonfederal lands, and were the DPS to be delisted, there would be very few mechanisms to curb ongoing habitat loss. On nonfederal lands since listing, over 300,000 of Murrelet habitat have been lost or anticipated for removal (Huff et al. 2006, p. 137). Nearly 100 percent of lost habitat on nonfederal lands since listing was due to timber harvest (Huff et al. 2006, p.126). From 1994 to 2004, 583,500 acres of Northern Spotted Owl habitat, which is similar to Marbled Murrelet habitat, were lost on non-federal forest lands (Raphael 2006).

 

Delisting would have dire consequences for Murrelets in the Northwest Forest Plan area, because it is estimated that over half of high-quality Murrelet nesting habitat in the NWFP area occurs on nonfederal lands (Huff et al. 2006, p. vii). Timber production priorities dominate forest management of most nonfederal forest lands in the Pacific Northwest.  In the NWFP area, 85% of the region's timber harvest between 2000 and 2005 occurred on private lands. Nonfederal forests are predominantly plantations that are managed in short rotations of 40–60 years, and the trend of private landowners to favor short rotations has actually intensified since the implementation of the NWFP due to the closure of mills capable of processing larger logs and the decline of Asian exports (McShane et al. 2004). On nonfederal lands, there are generally fewer protective measures for biodiversity or habitats such as riparian areas relative to federal lands, except for ESA-listed species (Suzuki and Olson 2008). Thus, were the Murrelet to be delisted, it would have few enforceable habitat protections on private lands in the Pacific Northwest.

 

In Oregon and Washington, there are no state laws or rules that explicitly require the   conservation of biodiversity on nonfederal lands. State forest practices rules do not generally address conservation or management of habitat at broad spatial scales, even though many industrial forests are large enough that landscape-level management guidelines could be considered. Moreover, some state forest practices rules may have unintended negative consequences on conservation by limiting landowners' options to enhance habitat heterogeneity (Suzuki and Olson 2008, p. 1034).

 

In Oregon, the relative threat of Murrelet habitat loss has not changed on non-federal land since listing. On private land in Oregon, little is currently required by State regulations in terms of murrelet habitat protection. There are 2,709,516 acres of commercial forest lands within the range of the murrelet in Oregon, much of which is privately owned and on which Murrelet habitat loss is ongoing. Even though the Murrelet is listed as threatened by the state of Oregon, this listing does not provide the species with any regulatory habitat protection (McShane et al. 2004, p. 4-108). Twelve-percent of habitat in the Oregon Coast Range is state-owned, but the state does not have regulatory mechanisms in place for Murrelet protection. State lands in Oregon are not being managed to increase the net amount of Marbled Murrelet habitat, and habitat loss is ongoing (McShane et al. 2004, p. 4-108, 4-13). The 5-year review states:

 

"The Service is aware of some occupied habitat on private lands in Oregon, but for the most part, the amount of suitable habitat on these lands is unknown. Due to the lack of State regulation of harvesting in murrelet habitat on private lands in Oregon and the failure to require pre-project surveys, it is likely that most suitable habitat has been lost from these lands over the 11 years since listing" (McShane et al. 2004, p. 4-4).

 

Despite Federal and State listing, FWS has allowed the Oregon Department of Forestry to harvest known occupied and known nesting sites. As of 2005, three planned and completed commercial timber operations occurred within Marbled Murrelet Management Areas on Oregon Department of Forestry lands in northwest Oregon (Smith 2005, p. 2). Logging in known Murrelet habitat includes both thinnings and clearcuts. The West Green Mountain clearcut was in an area not officially included in a Marbled Murrelet Management Area, but where subcanopy Murrelet behavior was detected (Smith 2005, p. 2). Clearcutting obviously destroys Murrelet habitat, but thinning is also detrimental. Potential effects of thinning include increasing population levels of predators in the stand, increasing predator access to nest cups, physical damage to nesting platforms, altered microhabitat conditions, and blowdown (Smith 2003, p. 3). Two thinned occupied sites blew down in winter 2007-2008, partially or wholly due to thinning operations. More harvests in known Murrelet habitat are planned. Given that Murrelet habitat is being logged despite current protections, the loss of habitat would be magnified even further were the tri-state population to be delisted. 

 

In California, occupied Murrelet habitat on non-federal land is protected under the California Forest Practice Rules and the California Endangered Species Act (CESA), under which the species is listed as endangered. California requires that timber harvest plans be reviewed by Department of Fish and Game biologists for impacts to Murrelets, and that landowners proposing harvest of suitable Murrelet habitat conduct surveys to demonstrate that the species is not occupying a stand prior to harvest. Despite these protections, Murrelet habitat loss in California is ongoing. A total of 5,933 acres of suitable Murrelet habitat were estimated to have been released for timber harvest on private lands in California not covered by a Habitat Conservation Plan (McShane et at 2004, p. 4-70).

 

In Washington, where the Marbled Murrelet is state-listed as threatened, the Department of Natural Resources Forest Practices Board adopted rules in 1997 to protect occupied Murrelet sites and non-surveyed Murrelet habitat within 50 miles of the coast that is not covered by a Habitat Conservation Plan. The State Environmental Policy Act has checklists that provide for the consideration of threatened species on private land, but this does not provide the species with any tangible protection. Timber, Fish, and Wildlife standards and guidelines protect riparian habitats that theoretically could protect some Murrelet habitat.

 

Marbled Murrelet habitat continues to be lost even with existing state and federal protections. Timber harvest allowed under Habitat Conservation Plans (HCPs) has

continued to eliminate habitat in California, Oregon, and Washington. Ownerships covered by HCPs account for the single largest area of Murrelet habitat loss (148,893 acres). Non-federal lands account for 80% of the total habitat anticipated for removal under Section 7 consultation (168,162 acres). Because some HCPs are long-term plans, all the acres of habitat removal consulted on may not have been harvested at present, but it is anticipated that harvest of the permitted acres will occur in the future. In Oregon, for example, the Elliott State Forest HCP will result in the loss of known occupied and suitable Murrelet habitat (ODF 2008).

 

Habitat Conservation Plans do not ensure the protection of Marbled Murrelets on covered lands for several reasons. Nonfederal landowners are not required to address the recovery of endangered species in an HCP. HCPs are only intended to maintain populations and habitats above baseline conditions, which are often determined by the initial population and habitat conditions upon which the agreement was signed. HCPs thus do not require landowners to improve habitat quality, increase populations, or create new habitats for listed species on their land. Further, most HCPs lack monitoring programs and many existing monitoring programs are insufficient to evaluate the HCP's success. Moreover, HCPs do not require landowners to modify management in response to new information because a ‘‘no surprise'' policy guarantees landowners that they would not be required to incur financial burden beyond the signed agreement (Suzuki and Olson 2008, p. 1027 - 1028).

 

In sum, the Marbled Murrelet continues to lose habitat in California, Oregon, and Washington, even with Endangered Species Act protection. Were the 3-state DPS to be delisted, there are no existing regulatory mechanisms which would adequately protect the species on federal, state, or private lands.

II. The range wide status of the Marbled Murrelet indicates that the species merits protection under the Endangered Species Act throughout its range.

As discussed above, the California, Oregon, and Washington Distinct Population Segment of the Marbled Murrelet merits uplisting to Endangered, and in addition to the lower 48 DPS, the species should be listed as threatened under the ESA in Alaska and British Columbia.

The International Union for the Conservation of Nature (IUCN) listed the Marbled Murrelet as Vulnerable in 2000 and changed the status to Endangered in 2004. The species was last assessed in 2008 and is still considered Endangered because "its population is estimated to have undergone a very rapid reduction over three generations (36 years), owing to a variety of threats" and the decline is "likely to continue" (IUCN 2008).

A. The Marbled Murrelet is declining sharply in Alaska and British Columbia.

The Marbled Murrelet has declined in Alaska by approximately 70 percent in the last 20-25 years (Piatt et al. 2007, p. 17). At-sea survey information spanning a wide geographic range indicates that the species declined annually by 5.4 – 12.7 percent at five of eight trend sites in Alaska and by 5.8 to 31 percent at eight of eight trend sites in British Columbia (Ibid.), where major population declines (22 to 73 percent) have been documented over a decade or more (Huff et al. 2006, p. 9). Piatt et al. (2007) state:

 

"Given that declines were estimated for sites over essentially the entire northern range of the species, there is cause for concern about the species' status . . . The declines are likely real, and related to combined and cumulative effects from climate-related changes in the marine ecosystem (most likely the 1977 regime shift) and human activities (logging, gillnet bycatch, oil pollution)" (p. 17-18)

In British Columbia the total population of Marbled Murrelets is currently estimated to be in the range of 54,300–92,600 birds (Piatt et al. 2007, p. 50). Piatt et al. (2007) conclude:

"All available evidence points to declining populations in British Columbia . . . [T]hese are high rates of decline for any population and indicate a potentially serious problem for the populations that they represent" (Piatt et al. 2007, p. 53).

Piatt et al. (2007) estimate that in the recent past, Marbled Murrelets in Alaska numbered on the order of one million birds (p. 1), and they describe the recent decline as "rapid and widespread" and "extraordinary" (pp. 49, 139). Because of the rate and magnitude of the decline, they questioned the accuracy of their analyses, but ultimately concluded that the decline was in fact real, range-wide, and biologically meaningful:

"The apparent change in population size and rates of decline reported here for murrelets are extraordinary and we therefore put them to the "straight-face test" by asking and answering four questions . . .

Are Declines of This Magnitude Even Possible? Yes . . . the observed rates of decline for Marbled Murrelets are not out of line with the range observed in other seabirds in Alaska. The difference, of course, is that murrelets are declining throughout their range and not just at a few sites here and there. . . we conclude that the trends observed in Alaska and British Columbia are possible from random variability alone, but more likely influenced to some degree by unnatural (anthropogenic) factors.

Is It Possible That the Methods or Statistics Have Led Us to Erroneous Measures of Trend? Yes . . In the end, however, we are struck by the consistency of the data in showing a declining trend, and apparent rates of decline ranging from -5.4 to -31 percent per year in the cases showing statistically significant change. Although the absolute rates could be inaccurate, the weight of statistical evidence leads us to err on the side of caution and assume that a biologically meaningful decline in the core population of murrelets has occurred during the past 15–20 years.

Are There Precedents for Changes of This Magnitude in Other Marine Wildlife Populations in the Northeastern Pacific Ocean? Yes. . . In particular, it appears that populations of closely related Kittlitz's Murrelet also are declining rapidly throughout their range (Kuletz and others, 2003; Robards and others, 2003), which overlaps almost completely with Marbled Murrelet from about Glacier Bay westward. Several species of seabirds in Alaska are undergoing similar or larger rates of decline at specific colonies (Dragoo and others, 2006), although no other species is currently undergoing declines in most areas of its range at the same time . . . There are parallels with changes in marine mammal populations as well (reviewed by Springer and others, 2007)" (Piatt et al. 2007, p.139-141).  

Other sources also corroborate the decline of Marbled Murrelets in Alaska and British Columbia. Klosiewski and Laing (1994) report a decline of 67-73% from 1972-1991; Lance et al. (2001) report a lack of recovery in Prince William Sound from 1989-1998 following the Exxon Valdez spill; Kelson et al. (1995) report a 40% decline in Clayoquot Sound, B.C., from 1982-1993; Kelson and Mather (1999) report a 22% decline in Clayoquot Sound from 1982-1996; Burger (1995b) reports a 50% decline in Barkley Sound from 1987-1993; Burger (2002a) reports an overall decline of 22-44% in British Columbia from1982-2002 (in Huff et al. 1996, p. 12).  

The National Audubon Society WatchList, an independent conservation status assessment of U.S. bird species, places the Marbled Murrelet in the Red category, which represents species considered to be declining rapidly, having very small populations or limited ranges, and facing major conservation threats (Piatt et al. 2007, p. 7).The Exxon Valdez Oil Trustee Council (2006) found that the Marbled Murrelet population in Prince William Sound is in decline, but that the relation of the decline to the oil spill is unclear, because the species is also declining in areas not affected by the spill. They suggest that other factors, including climate and prey shifts, are likely drivers (Integral Consulting, Inc., 2006 in Piatt et al. 2007, p. 7).

B. The Alaska and British Columbia Marbled Murrelet Population should be listed as threatened under the Endangered Species Act

1. The Alaska and British Columbia population of Marbled Murrelet is threatened by four of the listing factors under the Endangered Species Act.

A. The present or threatened destruction, modification, or curtailment of habitat or

range:

 

The dependence of Marbled Murrelets on old-growth forests for nesting habitat is well-documented (Ralph et al. 1995, Burger 2002, McShane et al. 2004, Piatt et al. 2007, p. 99). Marbled Murrelets in Alaska appear to use a broader range of habitats than populations at lower latitudes, and both tree and ground nesting occur in some areas, although evidence suggests that Murrelets prefer to nest in forested habitat (Kuletz et al. 1995a, Marks and Kuletz 2001, in Piatt et al. 2007, p. 99). In unforested areas where only ground-nesting opportunities are available, Murrelet densities are lower than in other areas (Piatt and Ford 1993).

 

Marbled Murrelet nesting habitat in Alaska is positively associated with large diameter trees with high epiphyte cover and large branches for nesting platforms. Average tree diameter in occupied areas is lower than in the contiguous 48 states, but nesting trees are typically the largest and oldest available (Piatt et al. 2007, p. 101, 103). Similarly, in British Columbia, Burger (2002) found that Murrelet nesting habitat is positively associated with large diameter old-growth trees, high density of large trees and nesting platforms, high vertical complexity in canopy structure, larger stands, and high habitat complexity (in Piatt et al. 2007, p. 104-105).

 

Because of their use of large, old-growth forest habitats for nesting, Marbled Murrelets in Alaska and British Columbia are threatened by habitat destruction, modification, and curtailment due to commercial timber harvest. Due to industrial-scale logging in the past half century, Marbled Murrelets have lost about 15 percent of their suitable nesting habitat in Southeast Alaska, and 33 to 49 percent in British Columbia (Piatt et al. 2007, p. 1). Piatt et al. (2007) state, "Evidence suggests that the past removal of large areas of old-growth nesting habitat in British Columbia and Alaska has had a direct negative impact on (Marbled Murrelet) populations" (p. 141). Future harvests and nesting habitat losses are likely to contribute to further reductions in Murrelet population size (Ibid.) The threat to Murrelet populations from logging is compounded by additional stressors the birds face in marine habitat (Ibid.). Further, because Murrelets are long-lived, there is likely a lag-time in population effects from past logging (Ibid.).

 

In addition to direct loss of nesting trees, Murrelets are also threatened by habitat fragmentation. In the lower 48, Murrelet nesting habitat is positively associated with low amounts of edge and fragmentation and large core areas of old-growth (McShane et al. 2004, p. 4-39). Murrelets appear to abandon highly fragmented areas over time (McShane et al. 2004, p. 6-6). The 5-year review states:

 

"[C]hanges in their (Marbled Murrelet) distribution and abundance have occurred in association with habitat loss and forest fragmentation (Service 1997). If murrelets are forced to utilize marginal habitat, nesting success could decline over time, leading to low nesting density and small populations (Raphael et al. 2002b). The fecundity rates of remaining pairs could also decline. Smaller patch size may also affect murrelet nest success and the number of nests, which may ultimately have long-term consequences on population size" (McShane et al. 2004, p. 4-109).

 

The negative consequences of fragmentation include effects on population viability and size, local or regional extinctions, displacement, fewer nesting attempts, failure to breed, reduced number of nests, lower nest success, increased predation, crowding in remaining patches and overall reduction of adult survival (Andrén 1995, Raphael et al. 2002b, McShane et al. 2004 in Piatt et al. 2007, p. 123).  Habitat fragmentation can cause population declines greater than those expected from habitat loss alone, due to edge effects which occur when predation rates are higher at habitat edges relative to interiors (Malt and Lank 2007, p. 160). Predation on Murrelets by Northern Goshawks could be greater in areas where both species are pushed into remaining old-growth patches (Piatt et al. 2007, p. 125).

 

The majority of studies have found that Murrelets avoid clear-cut edges and have lower nest success in more fragmented areas (Piatt et al. 2007, p. 105). Zharikov et al. (2006) found that nest success was lower near regenerating clearcuts, perhaps because of the regeneration of berry-producing shrubs which provide food for Murrelet nest predators. Steller's Jay is a key depredator of Murrelet nests, and several studies have shown that this Jay is most often found close to forest edges bordering clearcuts, and is more abundant in fragmented forest than in contiguous forest (Piatt et al. 2007, p. 123). Zharikov et al. (2006) also reported that nests were closer to clearcut edges than expected, but as Piatt et al. (2007) state, "[T]his cannot be interpreted that the murrelets had an affinity to forest edges or clearcuts, because the distances that were reported (mean distance to forest edge was 1.9 and 2.4 km at nest sites and randomly chosen points, respectively), were far greater than the distances that forest edge-effects are normally found (about 50–100 m)" (p. 105). Fragmentation also increases solar radiation and wind in the forest canopy which could affect the distribution of epiphytes, remove moss from nesting platforms, or cause overheating of eggs, chicks, or incubating adults (McShane et al. 2004, p. 6-6).

 

Some level of timber harvest has occurred throughout the entire old-growth forest nesting range of the Marbled Murrelet in Alaska and British Columbia, with industrial-scale logging occurring primarily in Southeast Alaska and British Columbia (Piatt et al. 2007, p.118). Twelve percent of old-growth forests, and 28 % of large-tree forest types have already been harvested in Southeast Alaska (Albert and Schoen 2006). In a comprehensive analysis of productive old growth stands in Southeast Alaska, Albert and Schoen (2006) found that from 1954 (when industrial logging began) to the present, there has been a 14-percent degradation of Murrelet habitat relative suitability scores for the region. By 2001, over 267,000 hectares in Southeast Alaska had been logged, including the clearcutting of 162,000 hectares of high-volume old-growth forest (Piatt et al. 2007, p. 119). Logging has been concentrated in highly-productive low elevation old-growth areas which also contain the forest types most preferred by the Marbled Murrelet (Piatt et al. 2007, p. 118).

 

Piatt et al. (2007) posit that the Tongass National Forest in Southeast Alaska may be the single most important forest habitat for Marbled Murrelets in the state as it encompasses the center of distribution for the species and the majority of the old growth western hemlock and Sitka spruce forests (p. 106). Over 184,856 ha of Tongass National Forest lands were cut by 1995, and an additional 303,750 ha were conveyed to the State of Alaska or to Alaska Native corporations, 60% of which were harvested by 1995. At least  364,500 ha of the Tongass have been harvested (DeGange 1996 in Piatt et al. 2007, p. 118). At a regional scale, harvests within eight regions in Southeast Alaska have removed 35–70 percent of the original large-tree forests (Piatt et al. 2007, p. 119). In addition to already lost and degraded habitat, about 30 percent of currently existing suitable habitat occurs on managed state, federal or private lands with potential for harvest (Piatt et al. 2007, p. 113).

 

Several independent studies show a significant reduction of nesting habitat due to industrial logging for Marbled Murrelet in British Columbia, ranging from a 33 to 49 percent reduction. The Ministry of Forests (2000) estimated that roughly half of Marbled Murrelet nesting habitat had been destroyed by 2000 (in Piatt et al. 2007, p. 119-123). Demarchi and Button (2001) found that logging, urbanization and agriculture have reduced overall habitat potential in British Columbia by 35% (in Piatt et al. 2007, p. 119-123). On Queen Charlotte Islands, Holt (2004) found that 42% of nesting habitat has been lost to industrial logging (in Piatt et al. 2007, p. 119-123). In the Sunshine Coast Forest District on southern mainland British Columbia, nearly 88% of Murrelet habitat has been lost (Piatt et al. 2007, p. 130-132).

 

Marbled Murrelet populations in British Columbia are likely to have declined by a similar proportion to the loss of forested habitat (Piatt et al. 2007, p. 123). The B.C. Marbled Murrelet population is likely to continue to decline as logging and to a far lesser extent road-building, urbanization and agriculture will continue to erode nesting habitat, and because the species also faces significant threats in the marine environment.

 

Throughout the forested regions of Alaska and British Columbia, loss of habitat from timber harvest will continue (Piatt et al. 2007, p.130). Further, suitable nesting habitat will not develop in young stands for 150 or more years and use of new habitat will lag for a considerable period because high philopatry leads to low rates of immigration (Albert and Schoen 2006, U.S. Forest Service 2002b in Piatt et al. 2007, p. 130).

 

In addition to logging, Murrelet habitat in Alaska is being lost due to the Spruce Bark Beetle infestation, and the death of Alaska yellow-cedars due to climate change (Piatt et al. 2007, p. 128). A Spruce Bark Beetle infestation recently destroyed mature spruce trees on over half of the forested land on the Kenai Peninsula and on 240,000 ha of land in the Copper River Basin (p. 114). The beetle has also caused a 22 percent reduction in Sitka spruce basal area on the Gulf Coast. It is not known how much Murrelet habitat has been lost due to the Spruce Bark Beetle, but is likely that Murrelet nesting habitat around Kachemak Bay was lost to beetle kill (Piatt et al. 2007, p. 128). The die off of Alaska yellow-cedar in Southeast Alaska and British Columbia, likely due to climate change, could also result in the loss of Murrelet nesting habitat.

 

Marbled Murrelet populations in Alaska and British Columbia are clearly threatened by the historical and ongoing loss of nesting habitat, which will continue into the future, and which will directly exacerbate the precipitous population decline facing the species in its northern range. The Alaska and British Columbia Murrelet population thus merits protection under the Endangered Species Act.

C. Disease or predation:

 

Marbled Murrelets are vulnerable to predation in all life stages, and populations of species that prey on Murrelet eggs, hatchlings, or adults are increasing. The 5-year review for the listed portion of the species states, "[I]t appears that the threat of predation on murrelets is greater than previously anticipated" (McShane et al. 2004, p. 6-29). Recent studies in the lower 48 have shown that the majority of Murrelet nests have failed and that the majority of known nest failures are due to predation (Ibid.). Piatt et al. (2007) analyzed Christmas Bird Count data and found that in British Columbia or Alaska, populations of five of seven known Murrelet predators have increased substantially (by 2–8 fold) during the past 30 years, including Bald Eagles (Haliaeetus leucocephalus), Common Ravens (Corvus corax), and Steller's Jays (Cyanocitta stelleri) (p. 143). They state:

 

"Although we have little quantitative data with which to assess the numerical impact of predation on murrelets, it is prudent not to underestimate the importance of this powerful force in regulating animal populations (Sinclair and others, 1998, Parrish and others, 2001, Springer and others, 2003). Predation is one of only a few factors we have identified that has the potential to cause the rates of change that we observed in murrelet populations" (p. 143).

 

An increase in the abundance of native predators can cause declines in avian prey populations (Cote and Sutherland 1997, Parrish et al. 2001 in Piatt et al. 2007, p. 126).

Marcstrom et al. (1989) found that an approximate doubling or tripling of the predator population, similar to the observed population increases of Murrelet predators in British Columbia, caused a rate of decline in prey as great as that observed for Marbled Murrelets in Alaska and British Columbia (Piatt et al. 2007, p.126).

 

In addition to mortality, predation can have multiple indirect effects on prey species including altered foraging behavior and altered timing of breeding (Piatt et al. 2007, p. 126). Predation pressure exacerbates the effects of habitat degradation, fragmentation, and destruction, and can directly slow or prevent population recovery (Ibid.).

 

Habitat fragmentation increases Murrelet vulnerability to nest predation. Nelson and Hamer (1995) found lower nesting success at sites located closer to forest edges and clearcuts. Burger et al. (2004) found that predation risk at Marbled Murrelet nests was higher near clearcuts and roads than in interior forests, and higher in fragmented landscapes than in relatively intact old-growth forests. Zharikov et al. (2006) found that nest success was lower near regenerating clearcuts, perhaps because of the regeneration of berry-producing shrubs which provide food for Murrelet nest predators, such as Steller's Jay. Steller's Jay is a key depredator of Murrelet nests, and several studies have shown that this Jay is most often found close to forest edges bordering clearcuts, and is more abundant in fragmented forest than in contiguous forest (Piatt et al. 2007, p. 123). Malt and Lank (2007) found that Murrelet nest disturbances by avian predators were significantly more frequent at hard edges relative to interiors, but less frequent at soft edges, and they did not find any edge effects at natural-edged sites (p. 160). The highest risk of nest predation has been documented in areas close to humans, likely because of food sources for predators such as corvids.  

 

In addition to increased levels of nest predation, the threat of high adult predation by raptors may have increased in some areas due to increased or recovering populations of falcons, eagles, and possibly some hawks. Predation on adult Murrelets likely has more negative consequences for population viability than take of chicks or eggs due to the importance of adult survivorship in demographic models (McShane et al. 2004, p. 6-29). 

Bald Eagles, Peregrine Falcons (Falco peregrinus), Northern Goshawks (Accipiter gentilis), Sharp-shinned Hawks (A. striatus), and Common Ravens are all known or suspected to prey upon adult Marbled Murrelets. Lewis et al. (2006) found that Marbled Murrelets are an important prey item for Northern Goshawks in Southeast Alaska. They recorded a total of 31 Marbled Murrelets delivered to 21 different goshawk nests. Murrelets sympatric with Northern Goshawks could experience greater predation if both species are forced into remnant old-growth patches (Piatt et al. 2007, p. 125).

 

Likely nest predators that occur commonly within the Alaska breeding range of Marbled Murrelets include Common Ravens, Steller's Jays, Northwestern Crows (C. caurinus), Sharp-shinned Hawks, and Great Horned Owls (Bubo virginianus). Potential mammalian nest predators in Alaska include forest mustelids (marten, Martes americana, and fisher, M. pennanti), red squirrels (Tamiasciurus hudsonicus), mice (Peromyscus spp.), and raccoons (Procyon lotor) (Nelson 1997, Raphael et al 2002b, Bradley and Marzluff 2003 in Piatt et al. 2007, p. 22).

 

There is a dearth of information on the incidence of disease in Marbled Murrelet populations in Alaska and British Columbia (Piatt et al. 2007, p. 22). For the listed portion of the range, the 5-year review states:

 

"Murrelets will likely be affected by one or more diseases or biotoxins in the near future because of the cumulative effects of stressors in both their marine and  forest environments. The potential for murrelets to contract West Nile Virus seems high given their nesting habits, the abundance of mosquitoes in forests, the presence of the disease in other forest-nesting bird species, and the proximity of many murrelet nesting areas to open fields, clearcuts, or areas of human activity where corvids are abundant. . . Recent emergence of bacterial, fungal, parasitic, and viral diseases and biotoxins in seabirds poses an increasing threat to Marbled Murrelets" (McShane et al. 2004, p. 3-68 - 3-72).

 

The 5-year review also asserts that the emergence of disease in seabirds could be an indicator of deteriorating ecological conditions:

 

"The recent emergence of diseases in free-ranging birds in coastal marine systems is an indicator of declining ecological integrity. Diseases in seabirds are expected to increase significantly in the near future as ecological stressors in the marine environment, primarily coastal pollution, increase. Combined with other environmental stressors such as ocean climate changes and habitat loss, diseases may be especially significant with respect to species with declining populations" (McShane et al. 2004, p. 6-34).

 

It seems likely that increased incidence of disease is a potential threat not only to the lower 48 Murrelet population, but also to populations in British Columbia and Alaska.

 

Predation pressure on the British Columbia and Alaska Marbled Murrelet population is clearly increasing and this pressure will continue to increase with ongoing habitat loss and fragmentation.  Predation, especially on Murrelet adults, is an important population driver. Because of the increased incidence of predation, the potentially increased incidence of disease, the known population declines, and the other threats facing the Marbled Murrelet in its northern range, it merits protection under the ESA in British Columbia and Alaska.

 

D. Other natural or human caused factors:

 

Other natural or human caused factors which threaten the Marbled Murrelet in British Columbia and Alaska include oil spills, incidental take in fisheries, reduced and altered prey availability due to a variety of factors including climate change and oceanographic variability, and disturbance from motor craft.

 

1. Oil-Spill Mortality

 

Marbled Murrelets are highly vulnerable to oil spills because they remain at sea most of their lives, forage by diving, and prefer near-shore habitats which coincide with high levels of vessel traffic (Piatt et al. 2007, p. 1). Oil pollution has been and remains a significant problem for Murrelets throughout their range (Carter and Kuletz 1995, McShane et al. 2004, Piatt et al. 2007). Murrelets are jeopardized both by large-scale catastrophic events, such as the 1989 Exxon Valdez oil spill which killed 10,000 – 22,000 Marbled Murrelets, and by chronic low-level exposure to oil pollution (McShane et al. 2004). It is possible that tens to hundreds of Murrelets are killed per year in Alaska and British Columbia by undocumented chronic oil pollution (Piatt et al. 2007, p. 73).

In the cold waters typical of Alaska and British Columbia, even small amounts of oil can cause mortality or metabolic problems for Murrelets (Ibid.). Not only crude oil, but also diesel fuel can ruin seabird plumage and disrupt thermoregulation (Ibid.).

 

There were approximately 1800 vessel-related spills and 2300 spills from other sources in coastal Alaska waters from 1995 – 2005 (Piatt et al. 2007, Fig. 17, p. 76). Spills from non-vessel sources include those related to oil-rig operations and transport activities (Alaska Department of Environmental Conservation, unpub. data, 2005 in Piatt et al. 2007, p. 76).

 

In addition to direct mortality, oil spills have other negative affects on seabird populations. High numbers of vessels concentrate in spill areas for response, clean-up, and monitoring efforts, which can displace Murrelets from preferred foraging areas (Carter and Kuletz 1995, Kuletz 1996 in Piatt et al. 2007, p. 74). Disruption in Murrelet distribution was observed following both the Exxon Valdez spill and the Windy Bay spill (Ibid.). Oil spills also cause immediate and long-term damage to forage fish populations.  

In Prince William Sound, herring and sand lance populations, which are key parts of the Murrelet diet, were impacted by the Exxon spill, and may still be experiencing population perturbations as a result (Ibid.). Herring are also a forage species for Murrelet in Alaska, and following the spill in Prince William Sound, herring did not spawn in historical locations and fish showed sub-lethal damage and larval malformations (Norcross et al. 2001). Fewer juvenile Murrelets fledged following the spill, likely due to both direct mortality effects and to impacts on forage fish populations (Piatt et al. 2007, p. 74).

 

In British Columbia, the risk of large, catastrophic spills in marine Murrelet habitat are high, and hundreds of small spills occur each year from fishing vessels, recreational craft, marinas, and fueling barges (Burger 1992 in Piatt et al. 2007, p. 77). Beach surveys conducted from 1989 to 1997 revealed that small quantities of oil were present on 8 percent of all surveyed areas in British Columbia and 17 percent of surveyed areas on west Vancouver Island (Burger 2002). Oiled birds composed over 12% of the carcasses detected on the surveys with reported causes of mortality, and this is "undoubtedly an underestimate of the actual number of oiled birds because evidence of oiling might be easily missed in carcasses that had been partly scavenged or were decomposed" (Ibid. in Piatt et al. 2007, p. 78). Piatt et al. (2007) state:

 

"A major concern in British Columbia is the relatively high frequency of chronic oiling on the west coast of Vancouver Island and possibly also for Haida Gwaii (Queen Charlotte Islands) (which was not adequately sampled; Burger, 2002). Both of these regions support high densities of Marbled Murrelets" (p. 78).

 

Although it is difficult to quantify the extent of Marbled Murrelet mortality due to oil-spills in Alaska and British Columbia, it is well documented that chronic oil spills are occurring and that this is causing mortality of Marbled Murrelets, which in conjunction with other threats in the marine and terrestrial environments, is contributing to the precipitous population decline the species is experiencing in its Northern Range. Piatt et al. (2007) state:

 

"We have evidence that oil pollution and gill-nets kill hundreds to thousands of Marbled Murrelets each year in Alaska and British Columbia . . . Recent analysis of a long-term dataset on survival of murres in the North Atlantic indicates that oil spills have a strong impact on over-winter survival of murres and that the effect is additive to effects of environmental variation (Votier et al. 2005). This could be important to populations under stress" (p. 142).

 

Large catastrophic oil spills also pose a significant threat to the Murrelet in British Columbia and Alaska. Piatt et al. (2007) state:

 

"The risks from a large catastrophic oil spill, similar to the Nestucca spill (Burger 1993a) or the EVOS (Piatt et al. 1990), are significant and if such a spill coincided with seasonal aggregations of murrelets it could cause a serious local population decline. Oil spills would kill a large proportion of adult birds, resulting in significant demographic impacts (Boulanger et al. 1999)" (p. 78).

 

Because of the threat of oil-spill mortality and other threats, the Marbled Murrelet deserves Endangered Species Act protection in British Columbia and Alaska.

 

2. Incidental Take in Fisheries

 

The incidental capture of birds in gillnets is a well-documented effect of fisheries on Marbled Murrelets (Piatt et al. 2007, p. 78). In the past, gill-net mortality has had significant impacts on some Murrelet populations (Piatt et al. 2007, p. 78). Reduced population size is the chief long-term impact to Murrelet populations from gill-net mortality. Reduced breeding success and loss of use of certain nesting locations could also occur. If only small numbers of birds nest in particular locations, then the loss of a few individuals could lead to loss of use of those nesting areas. McShane et al. (2004) state, "Regardless of the level of impacts, they are additive to other factors negatively affecting murrelet populations" (McShane et al. 2004, p. 5-24).

 

In Murrelet population simulation models, bycatch mortality has strong impacts because of the importance of adult survivorship in sustaining populations.  Because of this, and because high mortality can occur in specific areas, Piatt et al. (2007) caution, "Murrelets are highly vulnerable to gillnet fisheries" (p. 92).

 

In Alaska and British Columbia, Marbled Murrelets overlap with salmon gillnetting operations, especially in Prince William Sound and Southeast Alaska. Annual bycatch mortality is difficult to measure, but bycatch is thought to kill several thousand Marbled Murrelets per year (Piatt et al. 2007, p. 1). The salmon gillnetting season spans the Murrelet Breeding season, when peak numbers of birds are in near-shore waters. Fisheries that extend into the late season overlap with the Murrelet's pre-basic molt period, when the birds are flightless and particularly vulnerable (Piatt et al. 2007, p. 78).

Juvenile Murrelets are also particularly susceptible to gillnet mortality when fishing occurs later in the breeding season (Piatt et al. 2007, p. 88).

 

It is estimated that at least 500 Marbled Murrelets are taken annually in fisheries bycatch in Alaska in Prince William Sound, south Unimak, Cook Inlet, and Kodiak Island, and it is possible that 1,000 or more birds are actually taken (Piatt et al. 2007, p. 84). In the Prince William Sound-Copper River drift gillnet fishery, estimated incidental take of Brachyramphus murrelets in 1990 and 1991 ranges up to 2,000 birds (Ibid.). Piatt et al. (2007) state, "The size and geographic extent of Alaska's salmon fisheries could make bycatch a conservation issue" (Piatt et al. 2007, p. 86).

 

Over 15 types of fisheries operate in British Columbia, but the salmon fishery poses the greatest threat to Marbled Murrelets (Piatt et al. 2007, p. 89). When the Marbled Murrelet was listed as Threatened in British Columbia, bycatch in gillnets was cited as one of the major threats to the species, and has also been cited in recent reviews (Rodway 1990, Hull 1999, Burger 2002 in Piatt et al. 2007, p. 90). Smith and Morgan (2005) reported that Marbled Murrelets were known to be caught in at least three fisheries, and they estimated bird bycatch rates in British Columbia salmon gillnet fisheries to be between 0.019 and 0.219 birds per hour fished, or over 12,000 seabirds per year, including over 550 Marbled Murrelets (in Piatt et al. 2007, p. 92).

 

Marbled Murrelet mortality in gill-nets is known to be occurring in fisheries in British Columbia and Alaska, and adult mortality is known to have significant effects on population dynamics (Piatt et al. 2007).  In light of the dramatic population declines the Murrelet has experienced recently in its northern range, and the many other threats facing the species in the marine and terrestrial environment, the Murrelet should be listed as threatened in British Columbia and Alaska.

 

3. Effects of Sports Fishery on Marbled Murrelets

 

Sports-fishing can have negative effects on Marbled Murrelets due to injury from hooks and lines and also from disturbance caused by recreational boats in near-shore waters (Piatt et al. 2007, p. 93). Murrelets have been hooked on sports-fishing lures in several areas in British Columbia (Campbell 1967, Carter et al. 1995, Alan E. Burger, University of Victoria, personal observation, 2005 in Piatt et al. 2007, p. 93). Sports fishing areas and Murrelet concentrations overlap in several areas in British Columbia including Barkley, Clayoquot, and Desolation sounds, off the West Coast Trail, and off Haida Gwaii in the Queen Charlotte Islands.

 

4. Reduced and Altered Prey Availability due to Overfishing

 

Both Marbled Murrelet distribution and productivity is directly influenced by the quantity and quality of available prey (Peery et al. 2004b, Becker and Beissinger 2005 in Piatt et al. 2007, p. 71). Prey availability is a critical determinant of reproductive success for piscivorous seabirds, and forage fishes differ substantially in fat content, energy density, and quality (Anthony et al. 2000). There can be also be differences in energy density within the same fish species, depending on season, environmental conditions, sex, and/or physical condition (Robards et al 1999a, Anthony et al. 2000 in Piatt et al. 2007, p. 72). The nutritional quality of prey affects rates of delivery to chicks, chick growth, and chick survival (Golet et al. 2000, Suryan et al. 2002, Wanless et al. 2005 in Piatt et al. 2007, p. 72). For seabird chick growth, prey quality may be even more important than prey quantity (Kitaysky et al. 1999; Wanless et al. 2005 in Piatt et al. 2007, p. 72). Prey quality and quantity also affects parental condition and predation rates. If prey quality or quantity is low, parents must spend more time capturing and delivering prey, which is both energetically expensive and increases exposure to raptors and nest predators (Peery et al. 2004b, Piatt et al. 2007, p. 72).

 

The "junk food hypothesis," originally suggested as a potential explanation for the dramatic decline of Steller's sea lions (Eumetopias jubatus) in the Gulf of Alaska, is also relevant for Murrelet populations (Osterblom et al. 2008). According to the hypothesis, a dietary switch to prey of low energy content can have detrimental population-level effects. Osterblom et al. (2008) conclude that recent work on piscivorous seabirds provides substantial evidence indicating the relevance of this hypothesis in food webs. They posit that the system-wide shift in prey availability to low-quality prey could be coupled to large scale changes in climatological and oceanographic forcing, as well as to predation, overfishing, and/or competition (Osterblom et al. 2008).

 

In Alaska forage fish populations, including sand lance, capelin, and myctophids, are protected under the 1976 Magnuson-Stevens Fishery Conservation and Management Act, which protects some important prey species for Murrelets. Other important Murrelet prey species, however, which Murrelet consume as juvenile fish, are fished commercially as adults, including Pacific herring, walleye pollock, and Pacific cod (Piatt et al. 2007, p. 93). Pacific herring is one of the most important prey species for Marbled Murrelets in Alaska, and juvenile herring may be essential to successful chick rearing in some areas (Kuletz 2005). Although herring harvest is managed conservatively in Alaska, conflicts exist with other fisheries, they are preyed upon by a variety of marine species, and they are being negatively impacted by pulp mills, especially in Southeast Alaska. (Ibid.). The herring population in Prince William Sound is in decline, and the Marbled Murrelet population there has declined in tandem with the herring biomass (Kuletz 2005). In Prince William Sound, several important Murrelet prey species are in decline in addition to herring, which might be attributable to the mass release of hatchery-reared salmon beginning in 1977. Hatchery-reared salmon potentially compete with other forage fish that feed on zooplankton, such as juvenile herring, and juvenile and adult sand lance (Pearson et al. 1999, Purcell and Sturdevant 2001, in Piatt et al. 2007, p. 94). 

 

There is no restriction on the exploitation of forage species in British Columbia (Piatt et al. 2007, p. 93). Further, in British Columbia, Pacific herring are commercially exploited on a large scale (Ibid.). In addition to overfishing, Murrelet prey availability in British Columbia is negatively affected by salmon-pen aquaculture operations. Murrelets prey on euphausiids (krill), and euphausiids play an important role in the overall marine food web. Piatt et al. (2007) state:

 

 "Given the increase in salmon-pen aquaculture in British Columbia, and the support given to this industry by the provincial government, there are likely to be pressures to expand the krill fishery and increase quotas. An intensive euphausiid fishery might affect murrelets directly, because they eat large euphausiids, or it could disrupt the food webs on which they depend" (Piatt et al. 2007, p. 95).

 

The extent to which Marbled Murrelet populations are affected by reduced prey availability by fishing practices in British Columbia and Alaska is not known, but given the other threats facing the species and the known population decline, this threat should not be disregarded.

 

5. Reduced and Altered Prey Availability due to Marine Change

 

Large-scale shifts in marine climate contribute to biological regime shifts which affect the abundance and type of forage fish available for Marbled Murrelets and other seabirds (Anderson and Piatt 1999, Benson and Trites 2002, Chavez et al. 2003 in Piatt et al. 2007, p. 71). Most Marbled Murrelet prey species consume macro-zooplankton including copepods and euphausiids. Murrelets are thus linked to changes in ocean and climate conditions which influence zooplankton abundance, distribution, and species composition (Zebdi and Collie 1995, Williams and Quinn II 2000, Mackas et al. 1998, 2001, Boldt et al. 2005 in Piatt et al. 2007, p. 72). The El Niño-Southern Oscillation, Pacific Decadal Oscillation, and other factors that affect ocean temperatures and current flows impact prey distribution and alter Murrelet foraging locations and success rates (Ainley et al.1995 in Huff et al. 2006, p. 20).

 

In the Northeastern Pacific Ocean, ecological conditions cycle on a multi-decadal time scale between cold and warm phases. In the late 1970's, the Aleutian Low Pressure system shifted and intensified which led to stronger westerly winds and warmer surface waters in the Gulf of Alaska, and strengthened the flow of the Alaska Current. This regime shift strongly affected the food web in the Gulf of Alaska ecosystem (Anderson and Piatt 1999). Important Marbled Murrelet forage populations, such as pandalid shrimp, capelin, and herring, declined by 95 percent or more in less than 15 years. These taxa have yet to recover throughout much of their range. When capelin and herring collapsed, they disappeared from the Marbled Murrelet diet and were replaced largely by lean juvenile pollock (Piatt and Anderson 1996). Because pollock have substantially lower energetic value than fatty species like capelin, they are not assimilated as efficiently by growing chicks (Romano et al. 2006 in Piatt et al. 2007, p. 142). Because of its lower nutritional value, pollock consumption induces nutritional stress and results in little fat deposition in seabirds even when it is consumed at a caloric rate equal to more preferred species such as capelin (Ibid.) Piatt et al. (2007) conclude that the change in predator diets may have had a negative effect on productivity, recruitment and survival in Murrelets (p. 142).

 

Murrelet populations in Alaska began to show evidence of declines in the 1980's, following the regime shift (Piatt et al. 2007, p. 71). Murrelet diet is known to have been affected in Prince William Sound, along the Alaska Peninsula, around Kodiak Island, and in British Columbia (Piatt et al. 2997, p. 71). The changes in the marine environment are thought to be one of the factors leading to Murrelet declines (Ibid.). Further, changes in the North Pacific climate and ocean ecosystems may be increasing in frequency, which puts additional stress on the Marbled Murrelet (King 2005 in Piatt et al. 2007, p. 71).

Piatt et al. (2007) state:

 

"[C]hanges in marine climate and so-called "regime shifts" have clearly had pervasive impacts on marine communities in the Northeastern Pacific Ocean. Indeed, this is the only factor we have examined that has the potential to explain large declines in murrelet numbers over a broad area of the Northeastern Pacific Ocean" (p. 143).

 

Global climate change is likely exacerbating the effects of regime shifts. Following the 1977 regime shift, the Bering Sea did not return to cooler conditions concurrent with a shift to cooler waters further south, as was the case with previous shifts.  Based on this observation, PICES (2005) concluded:

 

"We hypothesize that the overall climate change occurring in the Arctic, as indicated by warmer atmospheric and oceanic temperatures and loss of 15% of sea ice and tundra area over the previous two decades is making the Bering Sea less sensitive to the intrinsic climate variability of the North Pacific. Indeed, when the waters off the west coast of the continental United States shifted to cooler conditions after 1998, the subarctic did not change (Victoria pattern), in contrast to three earlier PDO shifts in the twentieth century. Thus we project that the Bering Sea will more likely continue on its current warm trajectory. . ." (p. 40).

 

The combined effects of global climate change and natural regime shifts are likely contributing to the decline of Marbled Murrelets in their northern range. Piatt et al. (2007) conclude that combined and cumulative effects from climate-related changes in the marine ecosystem and human activities such as logging, gillnet bycatch, and oil pollution are responsible for the observed population decline of Marbled Murrelets in Alaska and British Columbia (p. 2).

 

Reduced prey availability due to marine change has direct repercussions for Marbled Murrelet populations and obviates the need for the protection of the species in the northern portion of its range under the Endangered Species Act, especially in light of the terrestrial threats which the species also faces in Alaska and British Columbia.

 

6. Vessel Disturbance

 

Marbled Murrelets forage in the near-shore environment, which makes them highly susceptible to disturbance from vessel traffic (Piatt et al. 2007, p. 1). Various studies have indicated that disturbance from motorized vessels can have negative affects on avian species including decreased reproductive success, behavior changes with consequent energetic impacts, distributional shifts and habitat displacement, and reduced access to optimal foraging areas (reviewed in Piatt et al. 2007, p. 95-96).

 

Several studies have indicated that Brachyramphus murrelets are displaced by vessel traffic, and possibly impacted energetically. Hamer and Thompson (1997) found that nearly 70% of Murrelets moved a short distance away when a vessel approached (in Piatt et al. 2007, p. 96-97). Kuletz et al. (2003) found that Kittlitz's Murrelets were temporarily displaced by boat traffic and that when boats were present fewer birds made foraging dives and more birds flew off the water. Hentze (2006) found that single birds were more likely to flush, and because single birds are more likely to be breeding individuals, suggested that boat disturbance disproportionately affects breeding birds (in Piatt et al. 2007, p. 96-97). Speckman et al. (2004) found that the approach of small boats caused Marbled Murrelets holding fish (presumably meant for their chicks) to dive and then swallow the fish. This behavioral response could be energetically costly for chick-rearing murrelets and could have implications for reproductive success. Vessel disturbance can increase energy demands on Marbled Murrelets and could lead to reductions in fitness under adverse environmental conditions (Agness 2005 in Piatt et al. 2007, p. 142).

 

Boat traffic is inevitable in many areas frequented by Murrelets, and the degree to which the species will habituate to this disturbance is unknown. Other species have shown only limited ability to habituate to disturbances (Burger and Gochfeld 1990, Bleich et al. 1994, in Piatt et al. 2007, p. 95). In recent years vessel traffic has increased in many remote areas of Alaska and British Columbia (Piatt et al. 2007, p. 95). Many areas in Alaska have relatively high vessel activity and overlap with relatively high murrelet densities including Kachemak Bay–lower Cook Inlet, areas of Kodiak Island, Kenai Fjords National Park and Preserve, Prince William Sound, Glacier Bay National Park, and areas adjacent to coastal cities in Southeast Alaska (Piatt et al 1997, p. 97).

 

In sum, the Marbled Murrelet clearly merits Endangered Species Act protection in Alaska and British Columbia due to other natural or anthropogenic factors including oil-spills, gill-net bycatch, and reduced and altered prey availability.

 

E. Inadequacy of existing regulatory mechanisms:

 

1. Alaska

 

The Marbled Murrelet is not adequately protected by existing regulatory mechanisms on federal, state, tribal, or private lands in Alaska. 

 

There have been no federal efforts or directives aimed specifically at conservation of the Marbled Murrelet in Alaska (Piatt et al. 2007, p. 132). In terms of federal status, the Murrelet is a "species of management concern" in FWS Region 7 because of declining populations, but this designation does not confer legal status with respect to funding of studies or requirements related to resource extraction or development (Ibid.).

 

The Marbled Murrelet occurs in the Chugach National Forest and the Tongass National Forest. There are no existing regulatory mechanisms that prohibit logging of Murrelet habitat in either of these forests. In the Chugach National Forest, Marbled Murrelets are a "management indicator species" and a "species of interest," but these designations do not confer the species with any regulatory habitat protection. In the Tongass, an interagency conservation assessment recommended that the National Forest consider a reserve-based approach to Murrelet habitat management, by targeting low-elevation old-growth forests within 31 mi of the ocean for preservation (DeGange 1996), but this approach has not been implemented. Some areas were surveyed by USFS during the 1990's, but these data have not been analyzed, and have not been systematically used to manage selection of timber sales (Piatt et al. 2007, p. 132).

 

In an analysis of change in Murrelet habitat suitability in Southeast Alaska since industrial scale logging began in the 1950's, Albert and Schoen (2006) found a 14-percent degradation of murrelet habitat relative suitability scores for the region. Further, about 30 percent of currently existing suitable habitat (based on the sum of habitat suitability index scores within the region) occurs on managed state, federal or private lands with potential for harvest (Piatt et al. 2007, p. 113). It is estimated that 175,400 - 188,800 ha of National Forest land have been harvested in Southeast Alaska (Barbour et al. 2005, Albert and Schoen 2006). In South-Central Alaska, it is estimated that 20.6 MMBF of timber were harvested on National Forest land from 1988-2001 (van Hees 2005). More than 47% of South-Central Alaska forest is classified as timberland (unreserved productive forest land). Of this 47%, state and local government owns 44%, 28% is private, and 23% is national forest (Piatt et al. 2007, p. 114).

 

The Tongass National Forest may be the single most important forest habitat for Marbled Murrelets as it encompasses the center of distribution for the species and the majority of the old growth western hemlock and Sitka spruce forests in the State (Piatt et al. 2007, p. 106). By 1995, 184,856 ha of Tongass National Forest lands had been harvested, and an additional 303,750 ha had been conveyed to the State of Alaska or to Alaska Native corporations, about 60 percent were harvested by 1995, bringing the total harvested forest estimate to 364,500 ha as of 1996. (DeGange 1996 in Piatt et al. 2007, p.118). The Tongass National Forest is currently on a 100-year timber rotation, which does not allow enough time for suitable nest-tree characteristics to develop for Marbled Murrelet breeding (DeGange 1996 in Piatt et al. 2007, p. 130).

 

In January 2008 the Record of Decision was signed for the new Tongass Land and Resource Management Plan. The plan allows for the loss of over 40% of original large-tree productive old growth in Southeast Alaska (USDA 2008, p. 29). The plan designates 3.5 million acres of the Tongass for development, including 775,000 acres for timber production. The plan places 445,000 acres of old growth into the suitable land base for harvest, and allows development in 2.3 million acres of Roadless Areas (USDA 2008). 

 

The most recent Land Management Plan for the Chugach was developed in 2002, and sets the allowable sale quantity of timber at 6.3 MMBF for 2002-2007, and then 10.6 MMBF for the following decade. In the Chugach National Forest, 35.4 MMBF were harvested from 1980 to 1999 at an average annual rate of 1.8 MMBF per year. On the Chugach National Forest on the Kenai Peninsula, the amount of mature and old-growth forest decreased by 3,110 ha (17 percent) over a 25-year period between 1974 and 1999 (Piatt et al. 2007, p. 119). Because the Chugach does not have a high percentage of classic large-tree Murrelet nesting habitat, logging could be disproportionately harmful for Murrelets.

 

State Lands

 

The Marbled Murrelet has no special status with the State of Alaska. Under the Alaska Department of Fish and Game (ADFG) Non-Game Conservation Strategy, the Murrelet is a "featured species." This designation guides funding of studies under the State Wildlife Grant program, but does not confer the species with any protective status.

 

Timber harvest on state lands is a direct threat to the Marbled Murrelet. There are no regulations prohibiting the logging of Murrelet nesting habitat on state lands. Approximately 15% of potential Murrelet habitat is owned by the state (areas within 60 km of shore in southeast, south-central, and southwest Alaska) (Piatt et al. 2007, p. 109). Barbour and others (2005) determined that by 2001, 12,400 ha of forest had been cut on State and other non-federal, non-tribal lands in Southeast Alaska. Over 303,000 ha of Tongass National Forest land had been conveyed to the State of Alaska or to Alaska Native corporations by 1995, at least 60% of which has been harvested (Piatt et al. 2007, p. 118). The State owns roughly 154,400 ha of land within the Chugach National Forest boundary (Piatt et al. 2007, p. 119).  In South-Central Alaska, it is estimated that 41.3 MMBF of timber were harvested on State land from 1988-2001 (van Hees 2005).

 

Private Lands

 

The Marbled Murrelet does not have any protection on private lands in Alaska.

Forests on private lands in Southeast Alaska are being clearcut resulting in increased forest fragmentation and general loss of suitable nesting habitat for Marbled Murrelets (Piatt et al. 2007, p.130). In Southeast Alaska, Albert and Schoen (2006) estimate that 129,500 ha of private or other non-Federal forested lands have been harvested. In South-Central Alaska, it is estimated that 2062.9 MMBF of timber were harvested from 1988-2001, and that 97 % of this harvest occurred on private land (van Hees 2005).

 

Native Lands

 

There are no regulations to protect Marbled Murrelet nesting habitat on Native lands in Alaska. Most timberland on native lands in Southeast Alaska has already been harvested (Piatt et al. 2007, p. 129). Barbour et al. (2005) determined that by 2001, over 79,600 ha of forest on Native Corporation lands had been harvested in Southeast Alaska. In South-Central Alaska, Native Corporations' hold 168,500 ha of land within the Chugach, and it is estimated that that several thousand acres of this have been harvested in the last 15 years (Piatt et al. 2007, p. 119).

 

Marine Habitat in Alaska

 

There are no Federal or State directives to conserve Marbled Murrelets or their foraging habitats at-sea in Alaska (Piatt et al. 2007, p. 133). In May 2000, Presidential Executive Order 13158 directed Federal agencies to work with States and other entities to develop a National system of Marine Protected Areas (MPAs). An ADFG Task Force was appointed in November 2001to develop a strategy for developing an MPA program for the State, but the task force was disbanded in 2003, and there is no further record of actions by the State on MPAs (Ibid.). The State of Alaska MPA definition is less stringent than the federal definition, and many areas considered by the state to be MPA's do not carry any restrictions (Ibid.).

 

It is clear that existing regulatory mechanisms are inadequate to protect Marbled Murrelets and their terrestrial and marine habitat in Alaska.

 

 

 

 

2. Canada

 

Federal

 

The Marbled Murrelet is considered a threatened species in Canada under the Species At Risk Act (SARA). SARA prohibits direct take of Murrelets, but affords the species with very little actual habitat protection.  SARA only requires that "No person shall damage or destroy the residence" of any threatened or endangered species.  A residence is defined under SARA as: "a dwelling-place, such as a den, nest or other similar area or place, that is occupied or habitually occupied by one or more individuals during all or part of their life cycles, including breeding, rearing, staging, wintering, feeding or hibernation," and has been interpreted to mean only the nest tree when it is occupied. Further, SARA only applies on federal lands, but most Murrelet habitat in British Columbia occurs on provincial lands. SARA includes provisions for the identification of critical habitat as part of a recovery plan, but only protects such habitat on federal lands. SARA allows outside groups to petition the Canadian Environment Minister to enact provisions to protect habitat, but the minister isn't required to do so and to date has never enacted additional protections. 

 

Under SARA, a Recovery Strategy sets out the overall strategy for recovering a listed species, and Recovery Action Plans provide the details and timing of recovery implementation. The Recovery Strategy for the Marbled Murrelet is still under governmental review and has yet to be implemented. Details of the Recovery Strategy, associated Recovery Action plans, and their implementation schedules are not yet available.

 

The Recovery Strategy is based on the 2002 Conservation Assessment of Marbled Murrelets in British Columbia. The goal of the recovery strategy is to limit the loss of Murrelet population and nesting habitat to less than 30% of 2002 levels by 2032, with the expectation that human-induced population decline will cease after 2032. This strategy is not adequate to protect the B.C. Marbled Murrelet, however, because there are no provisions that actually limit the loss of nesting habitat. Evidence of stand occupancy by Murrelets does not prevent the harvest of old-growth forests in British Columbia, either on private land or on public leased lands (Piatt et al. 2007, p. 130). Logging, road-building, urbanization and agriculture will continue to erode nesting habitat in British Columbia, and the projected loss of habitat is not known (Piatt et al. 2007, p. 130). Another issue is that the recovery goal of less than 30% loss is based on 2002 habitat levels, not historic levels.  In the Sunshine Coast Forest District, for example, from 50-70% of Murrelet habitat had already been lost by 1985, and by 2005 Murrelet habitat made up only 12% of the total forested area (Piatt et al. 2007, p. 132). Allowing another 30% to be lost would mean that only roughly 8% of original Murrelet habitat would remain in this district. Moreover, the strategy hasn't been enacted yet, and when enacted, has no regulations mandating the preservation of forest stands which support Marbled Murrelets. Federal regulatory mechanisms are thus inadequate to protect Marbled Murrelet nesting habitat in British Columbia.

 

Provincial

 

The majority (about 80 percent) of Murrelet nesting habitat in British Columbia falls under provincial jurisdiction. The Marbled Murrelet is a Red-listed Species in British Columbia, and is an Identified Wildlife Species under B.C.'s Forest and Range Practices Act (FRPA). FRPA calls for the development of an Identified Wildlife Management Strategy and for habitat protection in Wildlife Habitat Areas (WHAs). Wildlife Habitat Areas provide very little protection for the Murrelet because they are prohibited from impacting the timber base by more than 1%. This means that no more than 1% of productive mature forest that is suitable for harvest (e.g. doesn't occur on steep slopes, etc.), which comprises the bulk of remaining Murrelet habitat, can be protected.  This 1% cap applies to all identified wildlife species combined, not just the Murrelet, further limiting the availability of protected habitat specifically for the Murrelet (Piatt et al. 2007, p. 139). As of November 2006, 124 Marbled Murrelet WHAs had been established, which amounts to less than 2% of the habitat required to maintain 70 percent of the provincial population (Dechesne-Mansiere 2004). Piatt et al. (2007) state:

 

"In British Columbia, the current total area of protected habitat constitutes less than one-half of the 70 percent of 2002 habitat areas recommended for long-term maintenance by the Canadian Marbled Murrelet Recovery Team (CMMRT; Bertram et al. 2003). It is unclear whether the Identified Wildlife Management Strategy and other land-use planning exercises currently underway are sufficient to produce the outcomes called for in the CMMRT recovery strategy" (p. 129-130).

 

The Forest Practices Board, an independent public watchdog group in British Columbia that investigates how well industry and government meet the intent of British Columbia's forest practices legislation, released a report in 2004 concluding that conservation of Marbled Murrelet habitat under the Forest and Range Practices Act is limited and very slow (Ibid.).Since the Marbled Murrelet was listed as Threatened in 1990, about 400,000 ha of possible habitat has been logged, representing habitat that could likely support about 24,000 murrelets (Dechesne-Mansiere 2004). In March 2005, the Forest Board issued a strongly worded condemnation of the provincial government, stating that "there is a systemic failure in government policy to protect threatened species such as Marbled Murrelets on crown forest lands" (Forest Practices Board 2005 in Piatt et al. 2007, p. 137). 

 

Marine Habitat

 

Existing regulatory mechanisms are inadequate to protect Marbled Murrelets in their marine environment in British Columbia. Bycatch mortality due to entanglement in fishing nets remains a major threat to the Murrelet in British Columbia, and there has not been an intensive research effort to reduce bycatch (Piatt et al. 2007, p. 90, 92). There are no mechanisms to protect Murrelets from decreased prey availability due to overfishing, as there is no restriction on the exploitation of forage species in British Columbia (Piatt et al. 2007, p. 93).  

 

In sum, even though the Murrelet is listed as Threatened both federally and provincially in Canada, existing regulatory mechanisms do not effectively protect Murrelet nesting habitat or safeguard Murrelets in the marine environment.

 

Conclusion

 

In light of the drastic population declines the Marbled Murrelet has undergone in the northern portion of its range, and given the historical and ongoing loss of Murrelet habitat in British Columbia and Alaska, the ongoing threats to the species posed by gill-net mortality, reduced and altered prey availability, and marine climate change, the increased risks of predation and disease, and the inadequacy of existing regulatory mechanisms, the Marbled Murrelet clearly qualifies for protection as a threatened species under the Endangered Species Act in Alaska and British Columbia.

 

Thank you for the opportunity to comment.

 

On behalf of all co-signers,

                                   

 

Tierra R. Curry, M.S.                         D. Noah Greenwald, M.S.     

Conservation Biologist                      Biodiversity Program Director, Science Director

Center for Biological Diversity         Center for Biological Diversity

PO Box 11374

Portland, OR 97211

 

 

Literature Cited

 

Albert, D., and J. Schoen, J. 2006. Conservation assessment and resource synthesis for the coastal forests and mountain ecoregion in southeast Alaska and the Tongass National Forest, in Schoen, J. and Dovichin, E., eds., Southeastern Alaska and the Tongass National Forest: Conservation assessment and resource synthesis: Anchorage, Audubon Alaska and Nature Conservancy.

 

Bureau of Land Management. 2008. Final Environmental Impact Statement for the Revision of the Resource Management Plans of the Western Oregon Bureau of Land Management Districts. U.S. Department of the Interior. Available online: http://www.blm.gov/or/plans/wopr/final_eis/index.php

 

Dechesne-Mansiere, S.B.C., 2004. Marbled Murrelet habitat management in British Columbia: Background analysis report for the chair of the Forest Practices Board: Victoria, British Columbia, Forest Practices Board, 23 p.

 

Earthjustice. 2008. FOIA documents. Available online: http://www.earthjustice.org/library/references/murrelet-macdonald-foia-documents.pdf

 

Grantham, B.A., F. Chan, K.J. Nielsen, D.S. Fox, J.A. Barth, A. Huyer, J. Lubchenco, and B.A. Menge. 2004. Upwelling-driven nearshore hypoxia signals ecosystem and oceanographic changes in the northeast Pacific. Nature 429(6993): 749-754.

 

Huff, M.H., M.G. Raphael, S.L. Miller, S.K. Nelson, J. Baldwin, tech. coords. 2006. Northwest Forest Plan—The first 10 years (1994-2003): status and trends of populations and nesting habitat for the marbled murrelet. Gen. Tech. Rep. PNW-GTR-650. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 149 p.

 

IUCN. International Union for the Conservation of Nature. 2008. Red list of threatened species. Brachyramphus marmoratus.  http://www.iucnredlist.org/details/144286

 

Kuletz, K.J., 2005, Foraging behavior and productivity of a non-colonial seabird, the Marbled Murrelet (Brachyramphus marmoratus), relative to prey and habitat: Dissertation, British Columbia, University of Victoria.

 

Lewis, S.B., K. Titus, K., and M.R. Fuller. 2006. Northern goshawk diet during the nesting season in southeast Alaska. Journal of Wildlife Management 70: 1151-1160.

 

Malt, J. and D. Lank. 2007. Temporal dynamics of edge effects on nest predation risk

for the marbled murrelet. Biological Conservation 140: 160-173.

 

McShane, C., T. Hamer, H. Carter, G. Swartzman, V. Friesen, D. Ainley, R. Tressler,

K. Nelson, A. Burger, L. Spear, T. Mohagen, R. Martin, L. Henkel, K. Prindle,

C. Strong, and J. Keany. 2004. Evaluation report for the 5-year status review of the

marbled murrelet in Washington, Oregon, and California. Unpublished report.

EDAW, Inc. Seattle, Washington. Prepared for the U.S. Fish and Wildlife Service,

Region 1. Portland, Oregon.

 

Norcross, B.L., Brown, E.D., Foy, R.J., Frandsen, M., Gay, S.M., Kline, T.C. Jr., Mason, D.M., Patrick, E.V., Paul, A.J., and Stokesbury, K.D.E., 2001, Synthesis of life history and ecology of juvenile Pacific herring in Prince William Sound, Alaska: Fisheries Oceanography, v. 10, no. Suppl. 1, p. 42-57.

 

Oregon Department of Forestry (ODF). 1995. Elliot State Forest Habitat Conservation Plan. Final Plan May 1995. Available online: http://www.oregon.gov/ODF/STATE_FORESTS/esf-hcp.shtml

 

Oregon Department of Forestry (ODF). 2005. Final Rip Tide Pre-Operations Report. Astoria District. June 21, 2005. 

 

Oregon Department of Forestry (ODF). 2008. Draft Elliot State Forest Habitat Conservation Plan. August 2008. Available online: www.oregon.gov/ODF/STATE_FORESTS/elliott.shtml#Habitat_Conservation_Plan_

 

Osterblom, H., O. Olsson, T. Blenckner, and R.W. Furness. 2008. Junk food in marine ecosystems. Oikos 117(7): 967-977.

 

Piatt, J.F., Kuletz, K.J., Burger, A.E., Hatch, S.A., Friesen, V.L., Birt, T.P. , Arimitsu, M.L., Drew, G.S., Harding, A.M.A., and K.S. Bixler. 2007. Status review of the Marbled Murrelet (Brachyramphus marmoratus) in Alaska and British Columbia: U.S. Geological Survey Open-File Report 2006-1387, 258 p.

 

Piatt, J.F., and R.G. Ford. 1993. Distribution and abundance of Marbled Murrelets in Alaska. Condor (95)3:662-669.

 

PICES (Scientific Report No. 28). 2005. Report of the Study Group on Fisheries and Ecosystem Responses to Recent Regime Shifts. Jacquelynne R. King (Ed.). North Pacific Marine Science Organization (PICES), Sidney, B.C., Canada. Available online:

www.pices.int/publications/scientific_reports/Report28/Rep_28_FERRRS.pdf#page=7.

 

 

Rabot, T. 2007. Review of Service Decisions under the Endangered Species Act Involving the Deputy Assistant Secretary, Fish and Wildlife and Parks, in Region 1 (2001-2006). Memorandum to the Director of Fish and Wildlife Service. June 28, 2007. FWS/R1/AES.

Raphael, M.G. 2006. Conservation of listed species: the northern spotted owl and marbled murrelet. Chapter 7 in R.W. Haynes, B.T. Bormann, D.C. Lee, and J.R. Martin (technical editors), Northwest Forest Plan—the first 10 Years (1994–2003): synthesis of monitoring and research results. Gen. Tech. Rep. PNW-GTR. USDA Forest Service, Pacific Northwest Research Station, Portland, Oregon. http://www.fs.fed.us/pnw/publications/gtr651/ p 121

Smith, C. 2005. Preliminary Biological Assessment of the Rip Tide Timber Sale: Potential Impacts to the Tidewater Marbled Murrelet Management Area. Oregon Department of Forestry. February 2, 2005.  

 

Smith, C. 2003. Memorandum to T. Scoggins. Preliminary Biological Assessment of the Simmons Ridge Combination Timber Sale: Potential Impacts to the Simmons Ridge Marbled Murrelet Management Area. Oregon Department of Forestry. 14 January 2003. 

Smith, J.L., and Morgan, K.H. 2005.Assessment of seabird bycatch in longline and net fisheries in British Columbia: Delta, British Columbia, Canadian Wildlife Service, Pacific and Yukon Region, Technical Report 401.

Suzuki, N. and Olson, D.H. 2008. Options for biodiversity conservation in managed forest landscapes of multiple ownerships in Oregon and Washington, USA. Biodiversity and Conservation 17: 1017-1039.

U.S. Department of Agriculture. 2008. Tongass Land and Resource Management Plan. Final Environmental Impact Statement Plan Amendment Summary. Forest Service

R10-MB-603f. 35 pp.

 

U.S. Fish and Wildlife Service. 1997. Recovery Plan for the Threatened Marbled Murrelet (Brachyramphus marmoratus) in Washington, Oregon, and California. Portland, Oregon. 203 pp.

 

U.S. Fish and Wildlife Service. 2004. Marbled Murrelet 5-Year Review Process: Overview. August 31, 2004. 28 pp.

 

U.S. Fish and Wildlife Service. 2008. Endangered and Threatened Wildlife and Plants; 90-Day Finding on a Petition To Remove the California, Oregon, and Washington Population of the Marbled Murrelet (Brachyramphus marmoratus) From the List of

Endangered and Threatened Wildlife. Federal Register: October 2, 2008. Volume 73, Number 192. Pages 57314-57317.

 

U.S. Fish and Wildlife Service. 2008b. ESA Basics: More than 30 years of conserving endangered species. April 2008. available online: http://www.fws.gov/endangered

 

van Hees, W.S., 2005. Timber resource statistics of south-central Alaska, 2003: Portland, Oregon, U.S. Forest Service, Pacific Northwest Research Station, Resource Bulletin 248, 24 p.