29.10.2012 Views

Chapter 11 Science And Implementation

Chapter 11 Science And Implementation

Chapter 11 Science And Implementation

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 1<br />

<strong>Chapter</strong> <strong>11</strong><br />

<strong>Science</strong> and <strong>Implementation</strong><br />

Mary Ruckelshaus and Donna Darm<br />

The U.S. Endangered Species Act (ESA) relies heavily on science, and thus it is not surprising<br />

that science has become a major battleground in the controversy surrounding its implementation<br />

(Doremus this volume). It seems to have been the hope of Congress that decisions could be made<br />

based on science alone and that giving science a key role in decision making would insulate<br />

decisions from politics (H.R. CONF. REP. NO. 97-835, at 19). This hope was false from the<br />

beginning for at least two reasons. First, scientific information is almost always incomplete and<br />

uncertain. <strong>Science</strong> cannot answer with certainty many of the questions that must be answered in<br />

ESA decision making, in the time frames allowed by the statute. Second, while scientific data,<br />

analysis and advice should have a central role in informing natural resource decisions, scientific<br />

information by itself cannot “make decisions.” The Act’s seeming denial of the policy choices<br />

inherent in decision making, coupled with the difficulty of making hard policy choices that have<br />

regulatory effect, have led to criticism of the scientific process and decisions that result (Mapes<br />

2001; Boyle 2002; Dalton 2002; Seattle Times 2002; Stokstad 2002; Strassel 2002; Weber 2002;<br />

Cart 2003; Pianin 2003; Sacramento Bee 2003a & b; Cart and Weiss 2004).<br />

Congressional confidence in the ability of science to inform natural resource decisions is<br />

clear in the history of the ESA’s predecessors (Doremus 1997, this volume). With regard to<br />

listing determinations, for example, the Endangered Species Preservation Act of 1966 directs the<br />

Secretary to seek advice and recommendations from biologists and ecologists. The Endangered


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 2<br />

Species Conservation Act of 1969 was more forceful, directing the Secretary to make listing<br />

decisions on the basis of “the best scientific and commercial data available.” There is no<br />

definition of this term and the legislative history offers no clues. The requirement to use the best<br />

available science was incorporated into the 1973 Endangered Species Act, which was amended<br />

in 1982 (P.L. No. 97-304) to require that listing decisions under the ESA be based “solely” on<br />

the best scientific and commercial data available (ESA sec. 4(b)(1)(A)).<br />

The ESA requires agency reliance on science in other areas as well. The Secretary is to<br />

designate critical habitat based on the best scientific data available (ESA sec. 4(b)(2)). All<br />

federal agencies are to rely on the best available scientific and commercial data as they ensure<br />

their actions will not jeopardize the continued existence of listed species nor adversely modify<br />

their critical habitat (ESA sec. 7(a)(2)). Recovery plans are to adopt objective criteria for<br />

delisting (ESA sec. 4(f)(1)(B)(ii)). In requiring endangered species decisions to be made<br />

primarily on the basis of science, Congress sought to insulate the agencies from political<br />

pressure. Perversely, political pressure on the Services is intense, which in many cases has forced<br />

the policy choices inherent in science-based decisions underground (Doremus 1997).<br />

Some commentators have argued persuasively that “better science” will not settle the<br />

controversy but rather what is needed is more openness about the policy choices embedded in<br />

ESA decisions (Doremus 1997; Myer 2001; Yaffee 2005). The authors wholeheartedly agree, but<br />

we also believe better scientific information and better processes of eliciting and translating<br />

science can improve decision making under the ESA. The ESA has been the subject of intense<br />

debate in the scientific literature in terms of its effectiveness in protecting species (Schwartz<br />

1999; Boersma et al. 2001; Crouse et al. 2002; Scott et al. 2005). Furthermore, the question of


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 3<br />

how much scientific information is actually used in protecting species under the ESA recently<br />

was thoroughly evaluated (NRC 1995).<br />

This chapter focuses on where the ESA calls for science and examines agency practice in<br />

the use of science. We ask how well science is positioned to inform implementation of the Act<br />

and how well decision makers are prepared to make and explain decisions based on incomplete<br />

science (Doremus this volume). We draw on examples to illustrate our points, and it is only<br />

because of our greater familiarity with the National Marine Fisheries Service (NMFS) that we<br />

highlight more cases involving fish and marine mammals than terrestrial organisms. We close<br />

with suggestions for how scientists can better serve decision making under the ESA.<br />

Provisions of the ESA and Expectations of <strong>Science</strong><br />

The purpose of the ESA is to protect threatened and endangered species and the ecosystems on<br />

which they depend (ESA sec. 2(b)). To achieve that end, the Act requires the two agencies<br />

charged with implementation (the U.S. Fish and Wildlife Service [USFWS] and NMFS) to<br />

identify and list species that are threatened or endangered, and then provides substantive<br />

protections for listed species (table <strong>11</strong>.1). Once species are listed, the listing agency is to<br />

designate critical habitat and develop recovery plans for their conservation and recovery.<br />

In the sections that follow, we address the following questions for each stage of the ESA<br />

process from listing through recovery planning: (1) What is the role of science, and what has<br />

agency practice revealed to be the difficulties of incorporating science? (2) How have the public<br />

and courts responded? and (3) How could either the science, or the process of providing science,<br />

be improved?


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 4<br />

The <strong>Science</strong> Underlying Listing Determinations<br />

The listing process involves addressing two main biological questions: what is the species (or<br />

biological unit) to be listed? <strong>And</strong> what is the species’ likely risk of extinction? The USFWS and<br />

NMFS have developed interagency guidance on how to manage listing petitions, which primarily<br />

discusses how to navigate through the list of petitions for listing and for reconsideration of<br />

candidate species (USFWS and NMFS 1996a, 1999a), and does not offer biological criteria to<br />

address the issues below.<br />

What is a Listable Unit: Role of <strong>Science</strong> in Agency Practice<br />

The Act defines species to include “subspecies and any distinct population segment of any<br />

species of vertebrate fish or wildlife which interbreeds when mature” (ESA sec. 3(15)).<br />

The science job of identifying listable units involves analyses of reproductive isolation of the<br />

taxon in question. Taxonomic classification at the species level tends to be relatively<br />

straightforward, although it is rarely completed without challenges and revision. Evidence for<br />

reproductive isolation can come from several fronts, including molecular or allozymic genetic<br />

data, life history trait variation, estimates of intergroup dispersal, or geographic separation. Four<br />

difficult science issues have emerged in application of the ESA by USFWS and NMFS: Do all<br />

subspecies have equivalent “significance” taxonomically? What is a distinct population segment<br />

(Waples this volume)? How does hybridization between taxa affect species identification (Haig<br />

and Allendorf this volume)? How should artificially propagated individuals be considered?<br />

Species and Subspecies. The accuracy and degree of revisions of taxonomic classifications at the<br />

subspecies level varies greatly among species. The USFWS frequently has encountered<br />

situations in which it was uncertain whether a group of individuals should be classified as a


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 5<br />

distinct population segment, a subspecies, or even a separate species (e.g., interior least tern<br />

(USFWS 1985), lower Keys rice rat (USFWS 1991), Mississippi gopher frog (Rana capito<br />

sevosa), California red-legged frog (Rana aurora draytonii) (USFWS 1996a), and tiger<br />

salamander (Ambystoma californiense) (USFWS 2003c)). The listing determinations for the<br />

California gnatcatcher (Polioptila californica californica) (USFWS 1993; Zink et al. 2000),<br />

Dusky seaside sparrow (Ammodramus maritimus nigrescens) (Avise and Nelson 1989) and<br />

Florida panther (Puma concolor coryi) (Culver et al. 2000) aroused enormous controversy after<br />

the listings over the issue of whether the groups to be listed were significantly divergent enough<br />

from more common sister taxa to be considered subspecies, and therefore listable units. The<br />

social and political fallout from these listing decisions continues today (e.g., Carlson 2003;<br />

Miami Herald 2003; Pfeifer 2003; Wilson 2003). In some respects the controversy is misplaced,<br />

since even if they are not subspecies many of these population groups could likely be considered<br />

“distinct population segments” and so eligible for listing as a “species” under the Act (Easley et<br />

al. 2001). In other cases, new information about the lack of reproductive isolation may lead<br />

USFWS and NMFS biologists to conclude a group of populations is neither a subspecies<br />

(contrary to a published classification) nor a DPS, as in the case of the western sage grouse<br />

(USFWS 2004n).<br />

Distinct Population Segment. The USFWS and NMFS have a joint Distinct Population Segment<br />

(DPS) policy (USFWS and NMFS 1996b) that provides two tests of distinctness: (1) Is the<br />

population or group of populations markedly separate from other populations of the same<br />

species? and (2) is it significant? The policy provides a noninclusive list of factors that may be<br />

considered to qualify a population as significant, all of them biological (or scientific) factors.<br />

These include persistence in an ecological setting unusual or unique for the taxon, evidence that


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 6<br />

loss of the discrete population segment would result in a significant gap in the range of a taxon,<br />

evidence that the discrete population segment represents the only surviving natural occurrence of<br />

a taxon that may be more abundant elsewhere as an introduced population outside its historic<br />

range, or evidence that the discrete population segment differs markedly from other populations<br />

of the species in its genetic characteristics.<br />

Biologists may find it difficult to determine whether a population that is separate is also<br />

significant if they are uncertain of the taxonomic classifications above the population level.<br />

Making this determination can be very important for the degree of protection afforded a species.<br />

For example, in their first review of the status of southern resident killer whales (Orcinus orca),<br />

Federal biologists believed that the classification of orcas world-wide was probably wrong, but<br />

that was the taxon against which they felt they must determine “significance” (the second prong<br />

of the DPS test). In the absence of scientific consensus on how to group the units, the biologists<br />

initially concluded the southern resident group did not meet the DPS definition. Although the<br />

biologists making the determination did not dispute the declining trend and precarious status of<br />

the southern group of whales, because they could not be considered “significant” to the world-<br />

wide taxon, NMFS initially did not list the southern resident group of orcas under the ESA<br />

(NMFS 2002b). Upon reviewing new scientific information after a judge’s order to revisit the<br />

determination, the biologists concluded that the southern resident killer whales met both the<br />

discreteness and significance criteria for a DPS (Krahn et al. 2004). As a result, the southern<br />

resident DPS was proposed for listing as threatened under the ESA in December, 2004 (NMFS<br />

2004a).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 7<br />

NMFS adopted a policy for designating distinct population segments of Pacific salmon<br />

and steelhead prior to the joint DPS policy (NMFS 1991; Waples 1991, 1995). The NMFS policy<br />

relies on identification of evolutionarily significant units (ESUs). To be considered an ESU, a<br />

group of populations must be substantially reproductively isolated from other populations of the<br />

same species, and must represent an important component in the evolutionary legacy of the<br />

species.<br />

In delineating a DPS, one of the biggest challenges the two agencies face is how to<br />

address questions surrounding the significance of intraspecific variation in life histories, genetic,<br />

or morphological traits. It is important to determine the relationship of life history variants (e.g.,<br />

races) to one another in order to decide into how many pieces a species could or should be<br />

divided for listing determinations. For example, the unique life history of sockeye salmon<br />

(Oncorhynchus nerka), steelhead (O. mykiss), and chinook salmon (O. tshawytscha), which have<br />

both anadromous and nonanadromous (or resident) forms, pose an especially thorny problem. In<br />

many cases where the forms occur within drainages, the anadromous forms are at risk and the<br />

resident forms are not (NMFS 2003). The two life history forms are usually genetically more<br />

similar to one another than either is to its counterpart in nearby watersheds. Should each form be<br />

protected separately?<br />

Another example of difficulty in determining the significance of within-species variation<br />

is the USFWS listing of the California coastal gnatcatcher as a subspecies. The birds had<br />

originally been identified as a subspecies based on differences in bill size and shape, tail length,<br />

and overall coloration--all characteristics used by ornithologists to establish taxonomic<br />

classifications (USFWS 1993). These characteristics had evolved since the last ice age when the


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 8<br />

birds expanded northward from a refuge in Mexico--a very short time frame, evolutionarily<br />

speaking. Skeptical of the listing, private sector interests sponsored research which suggested<br />

that the morphological variations may not be genetically based. The USFWS revised its listing of<br />

the flycatcher, proposing to list it as a DPS instead (USFWS 2003). As a practical matter, in spite<br />

of a rash of new scientific information-gathering and analyses, the listing of the flycatcher as a<br />

DPS rather than a subspecies has little, if any, effect on the degree of protection it is afforded<br />

under the Act.<br />

These examples raise the problem of how to view the importance of morphological or<br />

behavioral variation that have evolved rapidly in response to dramatic environmental changes.<br />

How important is the variation to long-term persistence of the subspecies or species? Should<br />

these relatively newly evolved forms be protected? Some observers (and plaintiffs) argue that the<br />

Act is meant to protect morphologically unique forms as “distinct” population segments<br />

(Doremus 1997). Others suggest that the relevant inquiry should be whether, if lost, the variation<br />

could evolve again in a time span that would be meaningful to humans (such as a few<br />

generations) (Ruckelshaus et al. 2002a; Waples et al. 2004). Once a DPS is delineated, the<br />

relevant inquiry is whether the distinct population as a whole is at risk of extinction. This turns<br />

the difficult task back to the scientists of determining the importance of different morphological<br />

or life history forms to the continued existence of the distinct population segment (or species or<br />

subspecies) as a whole.<br />

Hybrids. Hybridization between closely related taxa can create listing challenges, especially<br />

when one of the hybridizing species is common and the other rare (Haig and Allendorf this<br />

volume) For example, the red wolf (Canis lupus rufus), which is listed as endangered under the


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 9<br />

ESA, may interbreed with the unlisted eastern gray wolf (Canis lupus lycaon) (Wayne and Jenks<br />

1991; Dowling et al. 1992; Nowak 1992; Wayne et al. 1992; Brownlow 1996). A less high<br />

profile but equally vexing example is the Westslope cutthroat trout (Oncorhynchus clarki lewisi),<br />

which hybridizes with introduced rainbow trout (O. mykiss) in the western United States<br />

(Allendorf and Leary 1988; Behnke 1992; Rubidge et al. 2001; Rubidge 2003; Taylor et al.<br />

2003). Application of a proposed policy on considering hybrids (or intercrosses) under the Act<br />

(USFWS 1996b) did not prevent legal challenge and extensive discussion of the treatment of<br />

hybrids in USFWS’s ultimate decision not to list the trout (USFWS 2003a).<br />

Artificial propagation. Both USFWS and NMFS have had to consider species identification<br />

questions for taxa that include artificially propagated individuals occurring in natural habitats<br />

(e.g., fish produced in a hatchery, captively bred birds). Until recently, both agencies uniformly<br />

judged the danger of extinction and the state of recovery based on naturally reproducing<br />

populations. Recently, however, NMFS has proposed a policy that would consider the risk of<br />

extinction of species based on the likelihood of extinction of the entire species, both artificially<br />

propagated and naturally produced components (NMFS 2004b, 2005). The proposal raises<br />

interesting policy and science questions. On the policy side familiar questions include the<br />

acceptable degree of uncertainty and risk, both with respect to biological issues (such as<br />

likelihood of persistence) and management issues (such as likelihood artificial propagation<br />

programs will continue to be funded at a certain level or operated in a certain manner). On the<br />

biological side, it is very difficult for scientists to incorporate artificial propagation into<br />

extinction risk models because data on breeding patterns, reproductive success, and movement of<br />

hatchery and wild fish are so poor. <strong>And</strong> on the interface between science and policy, the NMFS’<br />

proposed policy raises questions about the importance of a species’ “evolutionary trajectory.” Is


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 10<br />

a DPS that is in danger of diverging from a natural evolutionary trajectory because of artificial<br />

selection “in danger of extinction”? Presumably it could be, if the artificial selection makes it<br />

likely the DPS will no longer be significant to the taxon (or evolutionarily significant, in the case<br />

of an ESU; Myers et al. 2004).<br />

What is a Listable Unit: Public and Court Reaction<br />

Courts are generally unwilling to second-guess agency biologists when it comes to taxonomic<br />

classification or evaluation of extinction risk. The General Accounting Office reviewed 64 listing<br />

decisions by the USFWS between 1999 and 2002 and found that peer reviewers<br />

“overwhelmingly supported” the application of science to the decisions; and found further that<br />

courts overturned only two listing decisions because of improper use of scientific data (GAO<br />

2003). However, courts will intervene where judges believe NMFS or USFWS has failed to<br />

follow the statute, regulations, or policies, or where the judge believes the agencies have failed to<br />

adequately explain the connection between the data and the conclusion. For example, a district<br />

court invalidated NMFS’ decision to list naturally spawned but not the hatchery spawned Oregon<br />

coast coho, even though NMFS found them to comprise a single ESU (Alsea Valley Alliance v.<br />

Evans, [161 F. Supp. 2d <strong>11</strong>54, D. Oreg. 2001]). An appeals court threw out USFWS’s decision<br />

to list a population of the cactus ferruginous pygmy owl because USFWS failed to explain how<br />

the population was “significant” and therefore a DPS under the joint DPS policy (National<br />

Association of Homebuilders v. Norton, CV 00-0903 March 10, 2003). <strong>And</strong> a district court<br />

concluded NMFS did not use the best available science when it relied on a taxonomic<br />

classification for Orcinus orca that NMFS’ scientists considered out of date (Center for<br />

Biological Diversity v. Lohn 2003).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, <strong>11</strong><br />

What is a Listable Unit: Opportunities for Advancing the Role of <strong>Science</strong><br />

The National Academy of <strong>Science</strong> review of use of science in the ESA was supportive of the<br />

Evolutionary Unit concept (NRC 1995). Much of the ongoing scientific debate over DPS/ESU<br />

identification is over fairly technical points, such as how best to describe evolutionarily<br />

significant variation for protection (summarized in Ruckelshaus et al. 2002a; Waples this<br />

volume). Some observers feel the USFWS and NMFS have too narrowly defined distinct<br />

population segments on the basis of scientific criteria (Doremus 1997), arguing that the ESA was<br />

intended to protect populations that have aesthetic value, that are a keystone species within their<br />

ecosystem, or are in some other sense unique. The concern is that the rigid “scientific” approach<br />

embodied in the ESU and DPS policies ignores other values that are equally valid and that<br />

Congress intended to protect. The two agencies, in turn, have adopted policies intended to<br />

implement the congressional mandate that listing decisions be based on the best “scientific”<br />

information available. Moreover, the ESU and DPS policies in most cases provide workable<br />

guidance in determining whether a species exists for purposes of the ESA.<br />

Extinction Risk: Role of <strong>Science</strong> in Agency Practice<br />

Once the listable unit (i.e., species, subspecies, or DPS) is identified, its risk of extinction must<br />

be estimated under section 4(a) of the ESA (table <strong>11</strong>.1). The Act defines an “endangered<br />

species” as one that is “in danger of extinction throughout all or a significant portion of its<br />

range” (ESA sec. 3(6)) and a “threatened species” as one that is “likely to become an endangered<br />

species within the foreseeable future” (ESA sec. 3(19)). Risk evaluations are necessarily a<br />

combination of scientific analyses and policy judgments about the degree of “acceptable” risk


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 12<br />

and the time frames over which risk should be evaluated (Burgman 2005). Decision makers must<br />

then interject a judgment about whether a species’ risk of extinction triggers the statutory<br />

definitions of “endangered” or “threatened.”<br />

Qualitative approaches to estimating species risk, if transparent and systematic, can be just as<br />

reliable as quantitative approaches (e.g., Keith et al. 2004, McCarthy et al. 2004). Nevertheless,<br />

neither the USFWS nor NMFS regularly use widely accepted qualitative approaches to<br />

estimating extinction risk for listing decisions, such as the IUCN red-list criteria (IUCN 1994) or<br />

NatureServe’s heritage ranking system, originally developed by states’ heritage programs in<br />

collaboration with The Nature Conservancy (NatureServe 2003). NMFS implemented its own<br />

risk evaluation matrix to assess the status of over 50 ESUs of Pacific salmonids (Wainwright and<br />

Kope 1999), accounting for diversity and spatial distribution concerns, in addition to the more<br />

familiar abundance and population growth trend criteria present in other approaches (e.g.,<br />

Allendorf et al. 1997, Shelden et al. 2001). In the more recent listing determinations, NMFS<br />

incorporated a method of indicating the certainty of risk evaluations for each ESU, borrowing<br />

from a method of generating distributions of “certainty scores” used for the Forest Ecosystem<br />

Management Team (FEMAT 1993).<br />

Quantitative extinction risk models (known collectively as Population Viability Analyses, or<br />

PVA) require at a minimum information on population size, population growth rate, and<br />

variability in population growth rate over time (Dennis et al. 1991; Boyce 1992; Morris et al.<br />

1999). The critical first step of identifying demographically independent populations is almost<br />

never done in PVA analyses, in spite of evidence that ignoring population structure can cause<br />

grave errors in estimates of extinction risk (Morris et al. 1999). In a recent counter-example,


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 13<br />

NMFS identified independent populations before conducting viability modeling for Pacific<br />

salmonids (McElhany et al. 2000; Ruckelshaus et al. 2002b).<br />

The data needed to parameterize even the simplest PVA models are almost always incomplete<br />

(Reed et al. this volume). Furthermore, simple extensions of past trends are not always accurate.<br />

Additional uncertainties in using PVA to model extinction risk occur in decisions about model<br />

structure--for example, how to depict population responses at small sizes, the effects of density-<br />

dependent population regulation, and choice of a quasi-extinction threshold have huge effects on<br />

model results (Morris et al. 1999). Because of uncertain data inputs, it is important to explore the<br />

sensitivity of PVA results to alternative assumptions (e.g., Dennis et al. 1991; Holmes 2001;<br />

Holmes and Fagan 2002), as NMFS has done for estimating the status of Pacific salmon and<br />

Steller sea lions (Eunetopias jubatus) in listing and recovery decisions (Gerber and VanBlaricom<br />

2001; NMFS 2003; PSTRT 2002; WLCTRT 2003).<br />

Applications of PVA generally assume that the trends and variability in input parameters<br />

observed in the past are representative of those to be observed in the future period over which the<br />

population dynamics are projected. This is almost certainly an incorrect assumption--climate<br />

change, changes in human management of the landscape, introduction and spread of<br />

nonindigenous species, and changing rates and intensity of human-influenced catastrophes (e.g.,<br />

fire, toxic, or oil spills) are likely to result in novel future conditions. A relatively new and<br />

promising approach is to use scenario planning, whereby scientists ask whether an estimated risk<br />

of extinction (or any population outcome) changes under alternative views of future conditions<br />

(see “Recovery Planning” in this chapter; Clark et al. 2001; Carpenter 2002; Peterson et al.<br />

2003).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 14<br />

An emerging issue for the USFWS and NMFS is interpretation of the statutory definition of an<br />

endangered species as one that is in danger of extinction “throughout all or a significant portion<br />

of its range.” As mentioned previously, while early USFWS listings were often ambiguous about<br />

the basis for listing less than a full subspecies, recent practice has rested on the identification of a<br />

DPS. Dissatisfied with some determinations not to list, plaintiffs have begun to challenge the<br />

agencies for failure to separately examine whether a species, subspecies or DPS is in danger of<br />

extinction in only a portion of its range (flat-tailed horned lizard: Defenders of Wildlife v. Norton<br />

2001; Canada lynx (Lynx canadensis): Defenders of Wildlife v. Norton 2002; green sturgeon:<br />

Environmental Protection Information Center (EPIC) v. NMFS 2004.) The two agencies have so<br />

far not articulated an interpretation of this statutory phrase, but some implied interpretations<br />

could have important consequences for future listings. Recent decisions by the USFWS have<br />

applied a biological test that is similar to the significance test of the DPS policy--that is,<br />

examining whether a population group is biologically significant even though it is not discrete<br />

(e.g., Florida black bear: USFWS 1998; Canada lynx: USFWS 2000). It is unclear whether the<br />

USFWS and NMFS believe that a species in danger of extinction in only a portion of its range<br />

must be listed throughout its entire range (see Marbled Murrelet v. Lujan, No. C91-522WDR,<br />

1992 U.S. Dist. LEXIS 14645 (Sept. 16, 2992). If the test of significance is a biological one (for<br />

example, a species overall is neither threatened nor endangered, but is in danger of extirpation in<br />

a biologically important part of its range), the incongruous result from existing court decisions<br />

would be that a population group that is not a DPS but is in danger of extirpation could lead to<br />

the listing of an entire species that is neither threatened nor endangered.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 15<br />

Extinction Risk: Public and Court Reaction<br />

Doremus (1997) suggested that negative public reaction to listing decisions is a reaction to<br />

unbridled agency discretion in identifying species and determining risk of extinction. In general,<br />

however, it seems the public reaction to listing tends to be less about whether the USFWS and<br />

NMFS have misapplied science and more amazement that federal laws will protect such<br />

creatures as rats and bugs, often against private interests. Courts tend to be deferential to agency<br />

listing determinations (GAO 2003), except when they conclude that the agencies failed to follow<br />

the statute or agency regulations. Although public comment on listing proposals often contests<br />

the agencies’ analysis of extinction risk (e.g., recent northern spotted owl [Strix occidentalis<br />

caurina] and marbled murrelet [Brachyranphus marmoratus marmoratus] status reviews by a<br />

consulting firm contracted by USFWS; Daly 2003), the authors are unaware of any successful<br />

court challenges in that area.<br />

Extinction Risk: Opportunities for Advancing the Role of <strong>Science</strong><br />

The USFWS and NMFS recently completed an interagency policy outlining the criteria by which<br />

the they will evaluate the effects of federal, state, and local conservation efforts when making<br />

listing decisions (“Policy for Evaluating Conservation Efforts” USFWS and NMFS 2003). These<br />

so-called “conservation measures” are those that recently have been, or soon will be<br />

implemented, and whose effects on the risk status of the species under review have not yet been<br />

realized. The criteria include addressing how likely it is that specific conservation measures will<br />

be implemented, and if implemented, how likely they are to be effective. The latter question is a<br />

scientific one that in some cases is exceedingly challenging to address (see discussion in<br />

“Recovery Planning” below).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 16<br />

The USFWS and NMFS would be well served by adopting guidance for making endangered and<br />

threatened species determinations. The guidance should acknowledge that making listing<br />

determinations is not just a science exercise but has three important policy components: (1) the<br />

time period over which persistence should be measured, (2) the level of risk that results in a<br />

threatened or an endangered finding, and (3) the burden of proof for demonstrating the effects of<br />

conservation measures. Guidance needs to be flexible enough to account for the inaccuracy of<br />

extinction risk estimates and for the biological differences among different species. For example,<br />

the time period over which extinction is considered may depend upon the inherent variability in<br />

demographic characteristics of a species or the ability of scientists to forecast long-term trends.<br />

What level of risk results in an endangered or a threatened finding might also vary depending<br />

upon the uncertainties involved. The guidance would need to leave room for decision makers and<br />

scientists to work together to understand the biological implications of alternative risk levels<br />

(e.g., modelers can illuminate for decision makers what a 0.99, 0.95, or 0.80 probability of<br />

extinction looks like) so that the population status estimated is appropriate for the policy<br />

determination that must be made (Doremus this volume).<br />

Research is needed on how best to make population or species demographic parameter estimates<br />

from spotty census information (Reed et al. this volume). Abundance information for many<br />

species of conservation concern consists of presence/absence data, index counts, or censuses<br />

during a specific life stage that is easy to count, such as breeding aggregations. Making a<br />

determination about the viability status of a species requires that these population or species<br />

subsamples be translated into whole population or species counts--what are the best methods for<br />

making that translation? What are the advantages and pitfalls associated with different<br />

approaches to estimating species numbers from population subsamples?


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 17<br />

Additional important features that if incorporated could greatly improve quantitative extinction<br />

risk models for ESA are how factors in the environment might act together to accelerate or<br />

mitigate downward population trends, and how ecological interactions affect single-species<br />

dynamics (see also “Recovery Planning” below; rev. NRC 1995).<br />

The <strong>Science</strong> Underlying Critical Habitat Designations<br />

At the time of listing, the USFWS and NMFS are to designate critical habitat to the maximum<br />

extent prudent and determinable (see ‘4(b) Critical habitat designation’ in table <strong>11</strong>.1). If critical<br />

habitat is not determinable at the time of listing, they have one year to complete the designation.<br />

The Act defines critical habitat as “the specific areas within the geographical area occupied by<br />

the species . . . on which are found those physical or biological features . . . essential to the<br />

conservation of the species,” and specific areas outside the geographical area occupied by the<br />

species . . . upon a determination by the Secretary that such areas are essential for the<br />

conservation of the species.” From this construction, the statute seems to contemplate an<br />

approach to critical habitat designation that favors occupied areas: the agencies are first to<br />

identify habitat elements essential to species conservation (for example, a particular type of tree<br />

for nesting, vegetation for forage or cover, gravel streambeds for spawning, etc.) and then<br />

include in the designation any areas within the species’ present range where those elements are<br />

present. Only for areas outside the species’ present range must there be a determination that the<br />

area itself is “essential for conservation.” In practice, the agencies, plaintiffs and some courts<br />

have blurred the two standards and required that all areas, occupied or unoccupied, meet the test<br />

for unoccupied habitat: the area itself must be essential for conservation. For example, in a case<br />

involving the silvery minnow, the court stated that critical habitat “must be limited


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 18<br />

geographically to what is essential to the conservation of the threatened or endangered species.”<br />

(Middle Rio Grande Conservancy District v. Babbitt, 206 F.Supp.2d <strong>11</strong>56 (D.N.M. 2000)). <strong>And</strong><br />

in a case involving the Alameda whipsnake the court stated: “critical habitat for occupied land is<br />

defined in part . . . as specific areas ‘essential to the conservation of the species.’”<br />

(Homebuilders Association of Northern California v. USFWS, 268 F.Supp.2d <strong>11</strong>97 (E.D. Cal.<br />

2003)). .<br />

The USFWS and NMFS have long maintained that critical habitat designation adds little<br />

to species protection (Clark 1999; but see Taylor et al. 2005). Section 7 of the ESA requires<br />

federal agencies to ensure that their actions do not jeopardize species’ continued existence and<br />

do not destroy or adversely modify their critical habitat. The two agencies have for the most part<br />

treated an action that adversely modifies critical habitat as also jeopardizing the species’<br />

continued existence, making the prohibition against adverse modification redundant. Agency<br />

regulations defining both jeopardy and adverse modification in similar terms (actions affecting<br />

“both the survival and recovery” of the species) have reinforced this approach. Critics point out<br />

that critical habitat designation is especially important for species protection in unoccupied<br />

habitat, where the USFWS and NMFS may be less likely to reach a jeopardy finding (Taylor et<br />

al. 2003, 2005). Two separate reviews indirectly looked at the effects of critical habitat<br />

designations on reported trends in species abundance and content of recovery plans, and the<br />

results were mixed (Clark et al. 2002; Hoekstra et al. 2002; Taylor et al. 2003, 2005). The<br />

authors believe the current landscape is too unsettled to draw a reliable conclusion from past<br />

practice. Recent court decisions have invalidated the agencies’ regulatory definition of adverse<br />

modification as not being sufficiently tied to conservation (Gifford Pinchot Task Force v. U.S.<br />

Fish and Wildlife Service, No. 03-35279 (9th Cir. Aug. 6, 2004); Sierra Club v. U.S. Fish and


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 19<br />

Wildlife Service, 245 F.3d 434 (5th Cir. 2001). As future section 7 practice adjusts to the new<br />

legal rulings, the two tests may prove not to be redundant and the designation of critical habitat<br />

may indeed provide increased protection for listed species.<br />

An extension of the view that the designation of critical habitat has little or no impact<br />

argues that there are no economic consequences of designation to consider. The result is a history<br />

of critical habitat designations, where they have been made, lacking a meaningful analysis of the<br />

economic impact (See New Mexico Cattle Growers Association v. U.S. Fish and Wildlife<br />

Service, 248 F.3d 1277 (10th Cir. 2001). Successful court challenges to designations (or the lack<br />

of designations) have led to multiple requirements for the USFWS to designate habitat in very<br />

short time frames. Moreover, courts have ordered the agencies to consider economic impacts of<br />

designation, even if they are “coextensive” with the impacts of applying the section 7 jeopardy<br />

requirement (New Mexico Cattle Growers Association v. U.S. Fish and Wildlife Service, 248<br />

F.3d 1277 (10th Cir. 2001). This requirement is contrary to the best available science regarding<br />

economic analysis, which would require an estimate of the costs of designation based on a<br />

comparison of the world with and without the designation (OMB 2003), and the requirement to<br />

do a full economic analysis for a burgeoning number of court-ordered designations has severely<br />

strained USFWS’s resources.<br />

In response, the USFWS has vigorously objected to the requirement (e.g., testimony of<br />

Assistant Secretary Craig Manson 2003; 69 FR 3064, January 22, 2004). Past congressional<br />

efforts have failed to amend the ESA to change the timing of critical habitat designation to<br />

coincide with recovery planning instead of listing (See Chafee-Kempthorne bill, S. 105-<strong>11</strong>80, SS<br />

2(b) [1997]), but it remains a topic of congressional interest (see H.R. 2933 [2004]). For the


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 20<br />

moment, it appears unlikely there will be relief for the USFWS from the critical habitat<br />

requirement (Washington Post 2004).<br />

Role of <strong>Science</strong> in Agency Practice<br />

Section 4(b)(2) requires critical habitat designation to be based on the best scientific data<br />

available, although the USFWS and NMFS may exclude areas from designation if economic or<br />

other relevant impacts outweigh the benefits of designation. Under the straightforward<br />

interpretation of the language in the Act, the science task would also be fairly straightforward, at<br />

least with respect to occupied habitat. For many species there is a body of knowledge about the<br />

elements in their habitat that are essential for their persistence--what they eat, where they mate<br />

and nest, and so forth. Under the approach that requires a determination of essentiality for both<br />

occupied and unoccupied habitat, the task becomes much more difficult. Usually at the time of<br />

listing (and often many years after listing), the agencies know very little about species’ habitat<br />

needs and thus identification of critical habitat is highly uncertain. For scientists, the question is<br />

often not just how much occupied and unoccupied habitat the species needs, but how many<br />

populations must exist for the species, subspecies, or distinct population segment to be viable.<br />

Also, at the time of listing, there usually are limited data on land use patterns and economic<br />

activities and there is no experience to suggest how these activities would be modified as a result<br />

of section 7 consultations. This is information that is necessary for a scientifically meaningful<br />

economic analysis.<br />

The agencies’ joint designation of critical habitat for the Gulf sturgeon illustrates some of the<br />

issues surrounding such a designation. The two agencies examined the population structure and<br />

concluded that the seven extant populations were largely reproductively isolated, that


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 21<br />

maintaining genetic diversity was important, that therefore the populations at the extremes of the<br />

range are important to conserve (containing the extremes of diversity), that the intermediate<br />

populations are important for connectivity, and that therefore all habitat currently occupied by<br />

those seven populations is essential for conservation (USFWS 2003d). Like the judgments made<br />

in analyzing extinction risk, these are clearly not judgments made in a policy vacuum. The<br />

question of how many populations are needed for conservation and how much habitat each needs<br />

for conservation are not just scientific questions. The answers depend upon tolerance to risk and<br />

time scales over which the risk is considered.<br />

One of the more contentious debates surrounding critical habitat designation concerns the<br />

consideration by the USFWS and NMFS of economic costs of designation and their discretion<br />

under section 4(b)(2) to exclude areas from designation if the benefit of exclusion outweighs the<br />

benefit of designation. The two agencies have only recently begun to apply economic analysis in<br />

their designations, and their use of the science of economics is not well-developed. Their past<br />

practice of collapsing the jeopardy and adverse modification requirements into a single test has<br />

complicated the economic analysis.<br />

Public and Court Reaction<br />

The provisions for critical habitat designation have proven a major flash point for advocates and<br />

critics of species protection (e.g., Sacramento Bee 2003b; Wilson 2003; Cart and Weiss 2004). It<br />

is the most prominent place in the statute where habitat plays a role, consistent with the fact that<br />

Congress recognized in adopting the Act that habitat destruction was one of the leading causes of<br />

extinctions. It is also one of the few places in the statute where economics comes into play,<br />

making it a rallying point for the development-regulated community. Added to that is the


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 22<br />

perception of landowners that when private land is designated as critical habitat the federal<br />

government is “taking” their property and will restrict use of their land. Finally, the USFWS and<br />

NMFS have long maintained that critical habitat designation adds nothing to species protection<br />

and have resisted designating habitat altogether or have simply designated habitat without much<br />

analysis (Patlis 2001). Given these views and perceptions, it is not surprising that the agencies<br />

fail to make critical habitat designations, and critical habitat designations made frequently end up<br />

in court (GAO 2003).<br />

Courts have responded to the agencies’ treatment of critical habitat designation with impatience,<br />

chastising their reluctance to designate and ordering critical habitat designations to be completed<br />

expeditiously. That reluctance and impatience of the courts has created a cycle in which the two<br />

agencies lack adequate time to conduct a thorough analysis, and courts grant less deference than<br />

they might to a more thorough analysis. Further complicating implementation of this provision<br />

are court decisions finding the agencies’ regulatory definition of adverse modification invalid,<br />

cited above. The lack of guidance by the two agencies and a growing number of court opinions<br />

serve to make the situation more uncertain.<br />

Opportunities for Advancing the Role of <strong>Science</strong><br />

There are promising biological approaches that could be used to at least partially address the<br />

question of how much occupied and unoccupied habitat is needed for species persistence (Hanski<br />

1999), but the logistics of parameterizing analyses or models are daunting. For example,<br />

population matrix models can be used to address the question of how changes in survival at<br />

particular life stages affect overall population dynamics or persistence (rev. Caswell 2001). If the<br />

data are available, scientists could use such models to ask whether the quality or quantity of


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 23<br />

specific habitats affecting a species (at a particular life stage) significantly affects the overall<br />

population demography--in other words, the species’ likelihood of recovering. An example of<br />

such an application with a listed species is the endangered Kemp’s Ridley sea turtle<br />

(Lepidochelys kempii) , in which matrix models suggested that the survival of subadults and<br />

adults in the ocean was most critical to overall population status (Heppell et al. 1996; Heppell<br />

and Crowder 1998). Such a result could be used to help target critical life stages and the habitats<br />

upon which they depend. Another approach is to empirically relate (or predict with models)<br />

changes in habitat to changes in species status (e.g., Jenkins et al. 2003). Unfortunately, the<br />

answer to the last part of the question posed under section 4(b)(2)--what habitat conditions or<br />

amounts significantly affect life-stage-specific survivals?--usually is not known (see “Recovery<br />

Planning” below).<br />

The science of economics also could contribute to improving the designation process. Section<br />

4(b)(2) requires the USFWS and NMFS to consider the impacts of designation and balance the<br />

benefits of exclusion against the benefits of designation. Federal guidelines recommend putting<br />

the two types of benefits into the same metric in a cost-benefit framework (OMB 2003).<br />

Although information may be readily available that allows economic impacts to be quantified<br />

and monetized, quantifying the benefits to species from critical habitat designation is more<br />

difficult.<br />

Thus, best economics practice would have the agencies measure the incremental impact and<br />

benefit of designation, yet the courts have ruled otherwise (New Mexico Cattle Growers, supra.).<br />

How should the agencies proceed in this situation? Best economic practice would have them<br />

conduct a formal cost-benefit analysis, yet the short statutory time frames, limited information


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 24<br />

and resources, and considerable latitude for discretion suggest formal cost-benefit analysis may<br />

be neither possible nor necessary. One observer has suggested that approaches other than cost-<br />

benefit analysis, such as a cost-effectiveness framework, may be more appropriate (Sinden<br />

2004). This recommendation is consistent with OMB guidance in cases where benefits are<br />

difficult to monetize (such as benefits to health or the environment).<br />

The National Academy of <strong>Science</strong>s suggests changes to the critical habitat designation process<br />

that might require legislation, including the identification of habitat critical to survival at time of<br />

listing with the designation of the rest of critical habitat at the time of recovery planning (NRC<br />

1995). Tying critical habitat designation to recovery planning has many proponents. A bill<br />

introduced in the 103rd Congress would have amended the ESA to call for critical habitat<br />

designation at the time of recovery planning and mandated recovery plans within two years (S.<br />

105-<strong>11</strong>80 [1997]). The Department of the Interior has gone on record supporting such a<br />

connection (Manson 2004). The General Accounting Office recommends that the USFWS and<br />

NMFS adopt guidance on critical habitat designation (which probably would require a regulatory<br />

change [GAO 2003]).<br />

The authors agree with the General Accounting Office that it is within the agencies’ power to<br />

make some improvements in the process through strategic guidance. Clearer guidance would<br />

help scientists focus on the scope of their task. Until the agencies amend the regulatory definition<br />

of adverse modification, it will be unclear what standard the they are applying in their section 7<br />

consultations and whether they continue to view the prohibitions against jeopardy and adverse<br />

modification of critical habitat as providing redundant protection. Guidance on the economic<br />

analysis called for in the Act would also help the USFWS and NMFS expedite designations. In


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 25<br />

particular, criteria for determining whether consideration of economic or other relevant impacts<br />

outweigh the benefits of designation would be helpful.<br />

The <strong>Science</strong> Underlying Limitations on Federal Actions: Section 7 Consultation<br />

Federal agencies must ensure any action they fund, permit, or carry out does not jeopardize the<br />

continued existence of listed species or result in the destruction or adverse modification of their<br />

critical habitat (table <strong>11</strong>.1; section 7(a)(2)). When a federal agency intends an action that may<br />

affect a listed species, it must consult with the listing agency. For actions that adversely affect<br />

the species, the agency provides its biological opinion as to whether the action as proposed is<br />

likely to jeopardize the continued existence of a listed species or adversely modify its critical<br />

habitat. If the agency’s opinion is that the action is likely to cause jeopardy or adverse<br />

modification, it must offer a reasonable and prudent alternative. The consultation provisions of<br />

the statute require that all agencies “shall use the best scientific and commercial data available”<br />

in fulfilling the consultation requirement.<br />

Conducting an analysis of jeopardy and adverse modification is one of the most common tasks<br />

required of the USFWS and NMFS and at the same time one in which the standards are most<br />

obscure (Rohlf 1989, 2001). The statute does not define jeopardy or adverse modification. The<br />

two agencies have adopted regulatory definitions of these terms (USFWS and NMFS 1999b), but<br />

their consultation handbook lays out an analytical approach that does not track the regulatory<br />

definitions. The regulations define “jeopardize the continued existence of” to mean “to engage in<br />

an action that reasonably would be expected, directly or indirectly, to reduce appreciably the<br />

likelihood of both the survival and recovery of a listed species in the wild by reducing the<br />

reproduction, numbers, or distribution of that species.” Adverse modification is defined as an


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 26<br />

alteration that “appreciably diminishes the value of critical habitat for both the survival and<br />

recovery of a listed species.” This fairly straightforward test would seem to require the agencies<br />

to compare the likelihood of a species survival and recovery with and without the proposed<br />

action, or the value of critical habitat with and without the proposed action. If the likelihood of<br />

survival and recovery “with” the action represents a reduction that is “appreciable,” the action<br />

jeopardizes. If it appreciably diminishes the value of critical habitat for survival and recovery, it<br />

is an adverse modification.<br />

In practice, however, the agencies seldom take that approach, and the their consultation<br />

handbook lays out a different chain of logic. The handbook does not provide further guidance on<br />

what it means to appreciably reduce the likelihood of survival and recovery (USFWS and NMFS<br />

1999b). Instead, it directs the agencies to consider the status of the species, the environmental<br />

baseline, the effects of the action, and cumulative effects to determine whether the species is<br />

likely to survive and recover. The handbook does not say what the agencies should do after<br />

summing up those factors. If the species is not expected to survive and recover, how much must<br />

the action under consultation contribute to that failure before it is considered jeopardy? Or, less<br />

likely, if the species is expected to survive and recover, does it matter how much modification<br />

occurs to the species’ remaining habitat? Analysis of adverse modification of critical habitat is<br />

further complicated by two circuit courts invalidating the regulatory definition, as discussed<br />

previously.<br />

In addition to offering an opinion on whether an action is likely to jeopardize a species’<br />

continued existence or adversely modify its critical habitat, the USFWS and NMFS must issue<br />

an incidental take statement authorizing a given level of take associated with the proposed


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 27<br />

action. Where the action involves habitat modification (for example, a grazing allotment), the<br />

agencies must determine what level of take will be associated with the habitat modification.<br />

Although such a determination must be made based on scientific analyses, it is very difficult for<br />

scientists to quantify a species’ response to habitat alterations, especially smaller-scale changes<br />

in habitat.<br />

Although the statute and agency regulations require the use of the best available science by the<br />

federal agencies, this is an area in particular where the tools of science are inadequate, by<br />

themselves, to answer the questions asked. The agencies often lack information to determine in<br />

advance what effect an action is likely to have on the listed species. Another complication is that<br />

for many species, risks are cumulative, and assessing the effect of each individual action on<br />

species status is exceedingly difficult. A final confounding factor is the frequent reality that<br />

threats come from many different actions in different sectors, forcing the agencies to make a<br />

choice about how much of the conservation burden should fall on a given sector (Box <strong>11</strong>.1).<br />

Role of <strong>Science</strong> in Agency Practice<br />

The variety of agency action considered in section 7 consultations is extremely diverse, as are the<br />

species affected. This variety makes it challenging to draw useful lessons from agency practice.<br />

In Box <strong>11</strong>.1, we offer a few examples to help draw out some lessons for scientists and policy<br />

makers.<br />

Opportunities for Advancing the Role of <strong>Science</strong><br />

The main opportunity for advancing the role of science in implementing section 7 is for the<br />

agencies to provide clear guidance about the general standards for jeopardy and adverse


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 28<br />

modification. Action agencies frequently complain that “jeopardy” is ill-defined and<br />

implementation is not uniform. It has also been suggested that the USFWS and NMFS could<br />

provide clearer guidance for individual species, for example on identifying critically low<br />

population levels, viable population levels, and allowable levels of take.<br />

Any standards the USFWS and NMFS provide should allow for scientific information to be<br />

fully taken into account, along with the policy considerations. For example, how important a<br />

particular factor is in limiting species recovery should be relevant to any determination,<br />

especially when it is considered in the context of other factors limiting recovery. As with all of<br />

sections of the Act, a life-cycle framework for estimating the potential effects of an action on<br />

species status would appear to be the best way to adequately address the question posed in<br />

section 7. Whether that life-cycle framework takes the form of a quantitative, formal life cycle<br />

model or a qualitative framework within which the cumulative effects of actions throughout the<br />

life history are considered does not matter nearly as much as adopting that perspective. In<br />

general, because of the inherent scientific uncertainty in estimating the biological consequences<br />

of numerous, small-scale actions, section 7 consultations should be treated as experiments that<br />

are monitored and adjusted as needed over time (see Box <strong>11</strong>.1.)<br />

The <strong>Science</strong> Underlying Limitations on Private Actions<br />

Once species are listed as endangered, the Act prohibits any person from taking a member of that<br />

species (table <strong>11</strong>.1). Take is defined broadly to include harm, and harm can include destruction<br />

of habitat to the extent it actually injures or kills individual animals. <strong>Science</strong> comes into play<br />

when those whose actions may take a listed species seek an exception from the take prohibition.<br />

Exceptions can be granted under section 10 (habitat conservation plans or HCPs) or section 4(d).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 29<br />

Regardless of the legal avenue, the standard is similar--the proposed take cannot result in<br />

jeopardy to the species continued existence or the destruction of adverse modification of its<br />

habitat.<br />

Role of <strong>Science</strong> in Agency Practice<br />

The USFWS and NMFS face the same challenges in permitting take that they face in<br />

consultations with federal agencies. As in the case of consultations, they must consider the effect<br />

of a single action for which they may or may not have a larger context. Furthermore, the<br />

difference in time frames between consultations with federal actors and take permits for private<br />

actors is an added complication. Between federal agencies, the situation is often more fluid.<br />

Consultation can be readily reinitiated when there are changed circumstances or new<br />

information. Private parties, on the other hand, often make significant investments on the basis of<br />

the take permits they receive and accordingly seek a longer-term commitment from the federal<br />

agencies. In an effort to encourage more landowners to protect endangered species, the agencies<br />

adopted a series of policies to offer greater assurances that agreements with the federal<br />

government would be lasting, for example through the “No Surprises” (USFWS and NMFS<br />

1998) and “Safe Harbor” (USFWS 1999, 2001a & 2003b, Bean et al. 2001) rules.<br />

Opportunities for Advancing the Role of <strong>Science</strong><br />

The opportunities for improvement in the use of science under sections 10 or 4(d) are similar to<br />

those under section 7--if the USFWS and NMFS encourage transparent evaluation of the<br />

cumulative effects of actions, in light of the overall effect of other actions throughout a species’<br />

life cycle, the decisions under these sections of the Act will be much improved.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 30<br />

The Act poses a challenging but important question under sections 7 and 10 (as well as section<br />

4(f), recovery planning, discussed below) that requires description of at least three key biological<br />

linkages: (1) identification of landscape-level processes that drive environmental factors<br />

imperiling a species, (2) relationships between critical environmental factors and species status,<br />

and (3) effects of actions that can directly or indirectly affect species status through changes to<br />

environmental factors or the processes that control them. To really answer those questions<br />

requires years of data collection, observations, modeling, and then analysis for each species of<br />

concern. Undertaking such work is important so that a basic understanding of the causes of<br />

species declines can be developed.<br />

In the meantime, while we await more relevant results, decisions under the ESA must be<br />

made in the face of imperfect biological understanding. Biologically robust shortcuts (e.g.,<br />

identifying data or information critical to making such estimates and where the biggest errors in<br />

making such judgments are likely) are sorely needed to help inform decisions being made right<br />

now (e.g., Burgman 2005). Furthermore, a broad review of science underpinning HCPs found<br />

that scientific guidance underlying a monitoring plan and clearly stated monitoring requirements<br />

were lacking in most HCPs (Kareiva et al. 1998). Given the inability of science to assess the<br />

efficacy of actions described in HCPs, such agreements (and others made under section 7) would<br />

most wisely be treated truly as experiments and include clear provisions for monitoring.<br />

The <strong>Science</strong> Underlying Recovery Planning<br />

The Act requires the USFWS and NMFS to adopt recovery plans for listed species but does not<br />

specify a time within which plans must be completed (table <strong>11</strong>.1). The recovery plan is expected<br />

to describe the biological conditions that are necessary for recovery of the species, or the state


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 31<br />

under which the species can be delisted. Recovery plans do not have any regulatory effect, but<br />

they can be used to coordinate and guide the agencies’ decision making in section 7<br />

consultations or in issuing take permits across a species’ range. The Act has minimal<br />

requirements for recovery plans--they must specify objective measurable criteria for delisting,<br />

specific actions that will achieve those objectives, and an estimate of the time and cost involved<br />

in completing the actions.<br />

Role of <strong>Science</strong> in Agency Practice<br />

<strong>Science</strong> has a clear role to play in determining the objective, measurable criteria that will lead to<br />

delisting. The considerations are similar to those involved in the population viability analysis that<br />

starts the ESA process (although in practice once a species is listed there may be a higher<br />

threshold for determining it is recovered than for making a determination of “not warranted” in a<br />

listing context). Aside from describing the conditions that would give a high level of confidence<br />

in species viability, there also is a role for science in identifying factors limiting recovery and in<br />

determining the biological consequences of site-specific management actions aimed at<br />

recovering the species.<br />

The NMFS has a recovery planning guidance document, which is general and provides<br />

broad latitude in how plans are developed and in their content (NMFS 1992). In response to the<br />

need for more technical recovery planning guidance for the many listed ESUs of Pacific salmon,<br />

NMFS has written a conceptual document that describes what a viable, or delistable, salmon<br />

ESU looks like (McElhany et al. 2000). NMFS defines a viable salmonid population (i.e., a VSP)<br />

as one that has sufficient abundance, productivity or population growth rate, spatial structure,<br />

and diversity so that it has a negligible risk of extinction in 100 years. This construct has proven


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 32<br />

to be very useful in keeping the scientific basis for recovery plans consistent among the 26 listed<br />

ESUs of salmon for which recovery-planning analyses are underway (Ruckelshaus et al. 2002a).<br />

The fundamental technical questions that are addressed for salmon recovery planning are: (1)<br />

What was the historical population structure of the ESU? (2) What are the characteristics of a<br />

viable population for each of the historically independent populations in the ESU? (3) What are<br />

possible configurations (which might differ from historical conditions) of the distribution, risk<br />

status, and diversity characteristics of populations across a viable ESU? and (4) What suites of<br />

actions are needed for recovery of the ESU? Answers to questions 2 and 3 provide viability<br />

criteria for populations and ESUs, and analyses underlying question 4 allow for evaluation of<br />

alternative actions and their predicted effects on population and ESU status.<br />

As described previously in “Listing Determinations”, the NMFS uses a mix of<br />

quantitative extinction risk models and qualitative risk evaluations in setting viability criteria<br />

under recovery planning for Pacific salmon and many of its listed marine mammals and turtles.<br />

The USFWS approaches recovery planning differently than NMFS. Instead of establishing<br />

species-based viability criteria and then identifying which actions can achieve those criteria,<br />

USFWS focuses technical analyses in recovery planning on threats to species viability and<br />

actions to abate those threats. In a recent review of USFWS recovery plans, most reviewed plans<br />

described recovery criteria in qualitative, rather than quantitative, terms (Gerber and Hatch<br />

2002). The same review found that for those species whose status was improving since listing,<br />

more of their recovery plans contained quantitative criteria than for those listed species whose<br />

status was declining (Gerber and Hatch 2002).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 33<br />

The science questions in this implementation stage of the Act are very broad, so there<br />

isn’t likely to be a single answer that emerges from scientific analyses aimed at describing a<br />

recovered species or the actions that are sufficient to allow recovery. For example, for some<br />

species, there may be more than one way to achieve viability criteria, and if a species is<br />

comprised of a number of extant populations, there may be multiple configurations of population<br />

distribution and risk status that can constitute a recovered species or ESU (Ruckelshaus et al.<br />

2002b).<br />

NMFS is incorporating scenario planning into its estimates of the likely effects of habitat,<br />

hatchery, and harvest management actions on the population status of listed salmon<br />

(Ruckelshaus et al. 2002a). Alternative future conditions to be included in population risk<br />

models are being elicited from watershed councils so that they can have greater confidence that<br />

the set of actions they plan to implement will achieve the desired population status for the<br />

salmon returning to their streams.<br />

Opportunities for Advancing the Role of <strong>Science</strong><br />

An in-depth review of the science in recovery planning concluded that both the science<br />

underpinning recovery plans and how science is used in their implementation need improvement<br />

(Clark et al. 2002). The role of science at this stage is most appropriately to use analyses that<br />

illuminate consequences of alternative actions, rather than attempting to provide “the answer.” In<br />

such cases where alternative suites of actions to achieve recovery are possible, collaboration with<br />

policy and planning staff who will influence implementation of actions is important (Rinkevich<br />

and Leon 2000; Wollondeck and Yaffee 2000; Brick et al. 2001; Yaffee 2005). Furthermore, as<br />

with section 7 consultations and section 10 conservation plans, because of the uncertainty


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 34<br />

associated with predicting the effects of most actions on species status, implementation of<br />

actions should be treated as experiments that are monitored with vigilance (Boersma et al. 2001;<br />

Crouse et al. 2002).<br />

<strong>Science</strong> obviously can play a critical role in providing more relevant biological data<br />

pertaining to imperiled species. The importance of basic natural history information for<br />

informing decisions under the ESA cannot be overstated (e.g., What constitutes a reproductively<br />

isolated group of individuals for a given species? In which habitats does a species occur<br />

throughout its life cycle? What are survival rates for the species in alternative habitat types?<br />

What is the relative reproductive success of pairings between alternative life history types?)<br />

Scientists within and outside the USFWS and NMFS have called for greater attention to<br />

multispecies and ecosystem effects in recovery plans (USFWS and NMFS 1994; Miller 1996).<br />

Indeed, the agencies developed a joint policy on a desired “ecosystem approach” to the ESA<br />

(USFWS and NMFS 1994). The potential importance of such community and ecosystem-level<br />

effects to prospects for species recovery is large. Recent examples in the literature illustrating the<br />

potential importance of community-level effects to the status of listed species include north<br />

Pacific whaling effects on sea otters in Alaska (Springer et al. 2003), ecological functions<br />

provided by grizzly bears (Pyare and Berger 2003), and predation by Caspian terns (Sterna<br />

caspia) on juvenile salmon in the Columbia River (Roby et al. 2003). Nevertheless, how best to<br />

include such community- or ecosystem-level effects in a recovery plan is not clear. Some have<br />

cautioned that satisfying the desire for multispecies protection by including more species in a<br />

single recovery plan may not be the answer. Clark et al. (2002) conclude that multispecies plans<br />

may in fact reduce the focus on individual species to the detriment of their conservation status.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 35<br />

Finally, science-policy issues that could benefit from greater guidance from the USFWS<br />

and NMFS mentioned in previous sections of the Act also pertain to recovery planning. These<br />

include a clear discussion of what an “acceptable” risk might be under recovery planning for<br />

different species, transparent discussion of how uncertainty in biological conclusions was<br />

accounted for in decisions, and whether there are clear differences between jeopardy and<br />

recovery standards applied by the agencies.<br />

What Can Scientists Do to Improve ESA implementation?<br />

Originally, the ESA was formulated to address a significant biological fact--species extinctions<br />

were occurring at an alarming rate--thus biological information and analyses should be at the<br />

core of decisions pertaining to species protection and recovery. What have scientists contributed<br />

to these pressing biological questions? What can we do now to help redress the problems with<br />

how science is used in ESA decisions?<br />

The contributions of academic and agency science to addressing biological questions<br />

asked of ESA implementation have been unevenly distributed among topical areas. In particular,<br />

quantitative approaches and analyses designed to identify units for conservation and those used<br />

in estimating species viability (or conversely, risk of extinction) have received the lion’s share of<br />

attention in the scientific literature (fig. <strong>11</strong>.1). These methods are not without controversy, but<br />

they are relatively well tested and examined, and many of their limitations have been discussed<br />

(Waples this volume; Boyce 1992; Akçakaya et al. 1999; Morris et al. 1999; Coulson et al. 2001;<br />

Brook et al. 2000, 2002; Ruckelshaus et al. 2002a; Ellner and Fieburg 2003). Unfortunately, the<br />

quantitative approaches that are plentiful in the literature are useful for only a small fraction of


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 36<br />

the species considered under the auspices of the ESA, because of the dearth of data to<br />

parameterize even the simplest quantitative approaches, as mentioned above.<br />

The science of characterizing degrees of imperilment using qualitative approaches also<br />

has improved greatly in the 30 years since inception of the Act (e.g., IUCN 1994; Akçakaya et<br />

al. 2000; NatureServe 2003), and greater attention to these methods would be helpful in ESA<br />

decisions for a majority of the species considered (Keith et al. 2004; McCarthy et al. 2004).<br />

Analytical methods to address the remaining questions asked in the ESA implementation<br />

process have barely emerged in the scientific literature (fig. 1). In particular, the science<br />

underlying identification of the effects of actions on species status lags far behind. Such analyses<br />

are needed to address questions under sections 7 and 10 (i.e., do these actions significantly<br />

reduce the species’ likelihood of survival and recovery?), Section 4(b) (i.e., what habitat<br />

quantities and qualities are necessary for the survival and recovery of the species?), and Section<br />

4(f) (i.e., what actions are sufficient to achieve species viability criteria?). It is no wonder that<br />

analyses addressing these questions in the ESA are not prevalent in the literature--the work is<br />

challenging, time-consuming, and difficult to generalize across species and locations. Especially<br />

lacking are empirical studies or analytical approaches to addressing the effects of specific actions<br />

on particular life stages, and how those are manifest at the overall population-dynamic or species<br />

level.<br />

Communication between scientists and decision makers is critical so that scientific results<br />

are relevant and decision makers are ready to incorporate results in their deliberations. Simple in<br />

concept, this interaction is complicated in practice by the different worlds in which the two live--<br />

scientists can say “I don’t know” and can readily accept that it may take years to answer some


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 37<br />

scientific questions, while decision makers must make decisions in the limited time frames<br />

afforded by the Act. Those differences can lead to frustration on both sides. The authors believe<br />

from our own experiences that the effort is well worth the trials involved (Ruckelshaus et al.<br />

2002a). Previous studies have highlighted the need for help from conservation scientists in more<br />

effectively linking basic biology or ecology to management decisions (Floyd 2001; Clark et al.<br />

2002). A good example of this is the American Fisheries Society’s recent appointment of a<br />

special committee to explore the concept of “best available science” and what it might mean in<br />

practice for fisheries management (Harris 2003). Approaches such as those developed under<br />

decision theory (Clemen 1996; Burgman 2005) and multicriteria mapping (Arrow and Raynaud<br />

1986; Bana e Costa 1990) are potentially useful, but we could find no examples of how these<br />

tools have been applied for decision making under the ESA.<br />

As others have emphasized, science can have a significant impact on decisions under the<br />

ESA, as long as it isn’t relied upon to be the sole arbiter in decisions (Doremus 1997; see Yaffee<br />

2005). Scientists, for their part, need to clearly explain to policy and decision makers how their<br />

choices may be informed by science, and where those limits occur (e.g., how certain are<br />

predictions of a species persistence probability over the next 50 years? 100 years?) Furthermore,<br />

scientists and decision makers truly interested in clarifying the role that science plays in<br />

implementation of the ESA should be willing to participate in public forums, where the same<br />

transparency of data/inputs and assumptions in analyses can be openly discussed to illustrate<br />

how science is used in decisions under the Act. It is a rare manager of endangered species who<br />

will communicate through forums to which scientists are accustomed, such as the published<br />

literature (e.g., Rosenberg 2002). If scientists are free to interact with policy- and decision<br />

makers in other processes designed to encourage open exchange, a clearer understanding can be


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 38<br />

had of what questions are being asked, and which require science input or policy guidance, or<br />

both.<br />

It is critical that discussions between scientists and policy makers, and those involving<br />

the public, emphasize clearly what is the scientific basis for a result, and what additional policy<br />

determinations were brought to bear in making a decision under the Act. For example, all of the<br />

scientific results described above that are relevant to different questions under the Act<br />

incorporated one or more points of policy guidance, such as the acceptable risk of extinction, the<br />

time frame for evaluating risk or likelihood of recovery, or the acceptable uncertainty in the<br />

effects of actions on species status. To improve scientific credibility and agency decision-<br />

making, both scientists and decision makers need to be clear about where data end and analysis<br />

begins. Quantitative models often use a mix of actual data and assumptions to produce fairly<br />

precise results. The precision of the results often causes laypersons and decision makers alike to<br />

ascribe too much accuracy to model results. If scientists and decision makers do not clearly<br />

distinguish between facts and assumptions, and how each drives the results, laypersons will<br />

challenge the facts, rather than question the assumptions.<br />

A recent example of problems with data underlying ESA decisions involved the method<br />

of collecting samples of hair for DNA testing to detect the occurrence of the listed Canada lynx<br />

on U.S. national forests. The federal and state biologists charged with providing the hair samples<br />

for DNA testing were accused of tampering with the data, because they included hair samples of<br />

taxidermy and other lynx from outside the prescribed geographic sampling area in the mix with<br />

bobcat and hair samples of unknown origins. The biologists stated they included such lynx<br />

samples to verify the accuracy of the genetics laboratory conducting the testing; forestry industry


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 39<br />

representatives accused the biologists of trying to rig the results to reduce logging pressure on<br />

federal lands; and the genetics laboratory got caught in the cross fire (Dalton 2002; Mills 2002).<br />

A congressional investigation concluded that unauthorized samples were submitted in violation<br />

of the interagency sampling protocol (GAO 2002). The damage to the credibility of government<br />

science from this incident was discussed in papers throughout the country (e.g., Mapes 2001;<br />

Boyle 2002; Strassel 2002). Much of this controversy could have been avoided if the process of<br />

gathering and analyzing data and its interpretation was transparent.<br />

Finally, one of the most positive and lasting impacts scientists can have on improving the<br />

degree to which science informs implementation of the ESA is to push management experiments.<br />

Because the Act poses biological questions that almost always will have to be answered with<br />

imperfect information, the only way to encourage action in the face of uncertainty (and feel okay<br />

about it) is to encourage implementation of alternative actions as experiments. Furthermore,<br />

carefully estimating what we can learn from experiments before launching into controversial sets<br />

of actions is well worth the effort (e.g., Paulsen and Hinrichsen 2002), as is carefully monitoring<br />

the results. In the end, to enhance protection of species under the ESA, biologists must get<br />

involved. Such involvement is not without potential costs (e.g., Halpern and Wilson 2003), but<br />

conducting sound research is not enough to protect a species if the results from a beautiful<br />

biological study sit in a journal, unread.<br />

Acknowledgments<br />

We thank Krista Bartz for conducting the analyses for and producing Figure <strong>11</strong>.1. Michael<br />

Bean, Frank Davis, Mike Ford, Dale Goble, Jeff Hard, Linda Jones, Jim Lecky, Marta


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 40<br />

Nammack, Mark Plummer, Mike Sissenwine, Russ Strach, John Stein, and Usha Varanasi<br />

provided helpful comments on this chapter and greatly improved its message.<br />

Literature cited<br />

Ackakaya, HR, S Ferson, MA Burgman, DA Keith, GM Mace and CR Todd. 2000. Making<br />

consistent IUCN classifications under uncertainty. Conservation Biology 14: 1001-1013.<br />

Ackakaya, HR, MA Burgman and L Ginzburg. 1999. Applied population ecology. 2 nd Edition.<br />

Sinauer, Sunderland, MA. 285 p.<br />

Allendorf, FW ad RF Leary. 1988. Conservation and distribution of genetic variation in a<br />

polytypic species, the cutthroat trout. Conservation Biology 2: 170-184.<br />

Allendorf FW, Bayles D, Bottom DL, Currens KP, Frissell CA, Hankin D, Lichatowich JA,<br />

Nehlsen W, Trotter PC, Williams TH. 1997. Prioritizing Pacific salmon stocks for<br />

conservation. Conservation Biology <strong>11</strong> (1): 140-152.<br />

<strong>And</strong>reasen, L and C Springer. 2000. Hatcheries promote fish recovery. Endangered Species<br />

Bulletin 25: 32-33.<br />

Arrow, K, H Raynaud. 1986. Social Choice and Multicriterion Decision-Making. MIT Press,<br />

Cambridge, MA.<br />

Avise, JC and WS Nelson. 1989. Molecular genetic relationships of the extinct Dusky Seaside<br />

sparrow. <strong>Science</strong> 243: 646-648.<br />

Bailey, E. 2003. Judge orders new irrigation plan for Klamath, citing a threat to salmon;<br />

Wildlife officials and water regulators are ordered back to the drawing board. Los<br />

Angeles Times, CA. July 18, 2003.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 41<br />

Bana e Costa, C. Readings in Multiple Criteria Decision Aid. Springer, Berlin.<br />

Bean, M., J. P. Jenny and B. van Eerden. 2001. Safe harbor agreements. Conservation Biology<br />

in Practice 2: 8-16.<br />

Behnke, RJ. 1992. Native Trout of Western North America. American Fisheries Society<br />

Monograph, American Fisheries Society, Bethesda, MD.<br />

Boersma PD, Kareiva PM, Fagan WF, Clark JA, Hoekstra JM. 2001. How good are endangered<br />

species recovery plans? Bio<strong>Science</strong> 51:643-49<br />

Brook, BW, MA Burgman, HR Akcakaya, JJ O’Grady, and R. Frankham. 2002. Critiques of<br />

PVA ask the wrong questions: throwing the heuristic baby out with the numerical<br />

bathwater. Conservation Biology 16: 262-263.<br />

Brook, BW, JJ O’Grady, AP Chapman, MA Burgman, HR Akcakaya, and R Frankham. 2000.<br />

How accurate are the predictions of population viability analysis for endangered species<br />

conservation? Nature 404: 385-387.<br />

Brownlow, CA. 1996. Molecular taxonomy and the conservation of the red wolf and other<br />

endangered carnivores. Conservation Biology 10: 390-396.<br />

Boyce MS. 1992. Population viability analysis. Annual Review of Ecology and Systematics<br />

23: 481-506.<br />

Boyle B. 2002. Restoring credibility to government science. Seattle Times, Feb. 27<br />

Brick P, Snow D, Van De Wetering S. 2001. Across the Great Divide. Explorations in<br />

Collaborative Conservation and the American West. Washington, DC: Island Press


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 42<br />

Burgman, MA. 2001. Flaws in subjective assessments of ecological risks and means for<br />

correcting them. Australian Journal of Environmental Management 8: 219-226.<br />

Burgman, M. A. 2005. Risks and decisions for conservation and environmental management.<br />

Cambridge University Press, Cambridge, U.K..<br />

Carlson, C. 2003. Lawsuits planned to protect Florida panther. The Miami Herald, FL. April<br />

24, 2003.<br />

Carpenter, Stephen R. 2002. Ecological futures: building an ecology of the long now. Ecology<br />

83:2069-2083.<br />

Cart, J. 2003. Coalition fights plan to reopen dunes; Environmental groups decry policy reversal<br />

by BLM to allow access to off-road vehicles. Los Angeles Times, CA. July 4, 2003.<br />

Cart, J. and K. R. Weiss. 2004. Battle lines drawn on protection of species. Los Angeles Time,<br />

CA. December 5, 2004.<br />

Caswell, H. 2001. Matrix Population Models : Construction, Analysis, and Interpretation.<br />

Sinauer Associates, Sunderland, MA. 722 pp.<br />

Center for Biological Diversity v. Lohn, No. C02-2505L, 2003 WL 23004985 (W.D.Wash. Dec.<br />

17, 2003)<br />

Clark, J. R. 1999. Testimony of Jamie Rappaport Clark, Director, Fish and Wildlife Service,<br />

Department of the Interior, Before the Senate Committee on Environment and Public<br />

Works, Subcommittee on Fisheries, Wildlife, and Drinking Water, on S. <strong>11</strong>00, May 27,<br />

1999.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 43<br />

Clark JS, Carpenter SR, Barber M, Collins S, Dobson A, et al 2001. Ecological forecasts: An<br />

emerging imperative. <strong>Science</strong> 293:657-60<br />

Clark, JA, JM Hoekstra, PD Boersma. 2002. Improving U.S. Endangered Species Act recovery<br />

plans: Key findings and recommendations of the SCB recovery plan report.<br />

Conservation Biology 16: 1510-1519.<br />

Clemen, R. 1996. Making Hard Decisions: An Introduction to Decision Analysis. North<br />

Holland Series in System <strong>Science</strong> and Engineering. Volume 8, PWS Kent Publishing,<br />

Boston, MA.<br />

Cooperman, MS and DF Markle. 2003. The Endangered Species Act and the National Research<br />

Council’s interim judgment in Klamath Basin. Fisheries 28: 10-19.<br />

Coulson, T. G., M. Mace, E. Hudson and H. Possingham. 2001. The use and abuse of<br />

population viability analysis. TREE 16: 219-221.<br />

Crouse, DT, LA Mehrhoff, MJ Parkin, DR Elam, and LY Chen. 2002. Endangered species<br />

recovery and the SCB Study: A U.S. Fish and Wildlife Service perspective. Ecological<br />

Applications 12: 719-723.<br />

Culver, M, WE Johnson, J Pecon-Slattery, and SJ O’Brien. 2000. Genomic ancestry of the<br />

American puma (Puma concolor). Journal of Heredity 91: 186-197.<br />

Dalton R. 2002. Fur flies over lynx survey’s suspect samples. Nature 415:107<br />

Daly, M. 2003. 2 private firms to review status of owl, murrelet. Seattle Times, 30 September,<br />

2003.<br />

Defenders of Wildlife v. Norton, 2001. 258 F.3d <strong>11</strong>36 (9th Cir.2001)


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 44<br />

Defenders of Wildlife v. Norton, 2002. 239 F.Supp.2d 9 (D.D.C. 2002)<br />

Dennis, B., P. L. Munholland and J. M. Scott. 1991. Estimation of growth and extinction<br />

parameters for endangered species. Ecological Monographs 61: <strong>11</strong>5-143.<br />

Doremus, H. 1997. Listing Decisions under the Endangered Species Act: Why Better <strong>Science</strong><br />

Isn’t Always Better Policy, 75 Washington University Law Quarterly 1029.<br />

Dowling, TE et al. 1992. Response to Wayne, Nowak and Phillips and Henry: use of molecular<br />

characters in biology. Conservation Biology 6: 600-603.<br />

Easley et al. 2001. THE ENDANGERED SPECIES ACT, The Stanford Environmental Law<br />

Society, Editors-in-Chief Easley et al.<br />

Ellner, Stephen P., Fieberg, John. 2003: Using PVA for management despite uncertainty: Effects<br />

of habitat, hatcheries, and harvest on salmon. Ecology: Vol. 84, No. 6, pp. 1359–1369.<br />

Environmental Protection Information Center (EPIC) v. NMFS, 2004. No. C-02-5401 (N.D.<br />

Cal. filed Mar. 2, 2004).<br />

Floyd, T. 2001. Complexity simplified (but who’s paying attention?) Ecology 82: 904-905.<br />

Forest Ecosystem Management Team (FEMAT). 1993. Forest ecosystem management: an<br />

ecological, economic and social assessment. Washington, DC: U.S. Forest Service and<br />

collaborating agencies.<br />

U.S. Fish and Wildlife Service (FWS). 1984. Final Decision to List the United States<br />

Population of Wood Stork, 49 FR 7332 (Feb. 28, 1984).<br />

U.S. Fish and Wildlife Service (FWS). 1985. Interior Population of Least Tern to be<br />

Endangered, 50 FR 21784 (May 28, 1985)


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 45<br />

U.S. Fish and Wildlife Service (FWS). 1986. Final Decision to List the Florida Breeding<br />

Population of Carcara, 51 FR 22838 (June 23, 1986);<br />

U.S. Fish and Wildlife Service (FWS). 1991. Endangered Status for the Lower Keys Population<br />

of the Rice Rat (Silver Rice Rat), 56 FR 19809 (April 30, 1991)<br />

U.S. Fish and Wildlife Service (FWS). 1993. Determination of Threatened Status for the<br />

Coastal California Gnatcatcher, 58 FR 16742 (March 30, 1993)<br />

U.S. Fish and Wildlife Service (USFWS). 1996a. Endangered and threatened wildlife and<br />

plants; determination of threatened status for the California Red-legged frog. 50 CFR<br />

Part 17 Vol. 61, No. 101. Thursday, May 23, 1996.<br />

U.S. Fish and Wildlife Service (FWS). 1996b. Draft Intercross Policy, 61 FR 4710 (February 7,<br />

1996)<br />

U.S. Fish and Wildlife Service (FWS). 1997. Cactus ferruginous pygmy owl. 62 Fed. Reg.<br />

10,730 (Mar. 10, 1997).<br />

U.S. Fish and Wildlife Service (FWS). 1998. Endangered and Threatened Wildlife and Plants;<br />

New 12-month finding for a petition to list the Florida Black Bear. 50 CFR Part 17. 63<br />

Fed. Reg. 67613 (Dec. 8, 1998).<br />

U. S. Fish and Wildlife Service (FWS). 1999. Safe Harbor Agreements and Candidate<br />

Conservation Agreements with Assurances. Final Rule with correction. Federal<br />

Register: Volume 64, Number 189; Page 52676.<br />

U.S. Fish and Wildlife Service (FWS). 2000. Endangered and threatened wildlife and plants;<br />

determination of threatened status for the contiguous U.S. distinct population segment of<br />

the Canada lynx and related rule; final rule. 50 CFR Part 17; 68 Fed. Reg. 40076 (March<br />

24, 2000).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 46<br />

U. S. Fish and Wildlife Service (FWS). 2001a. Response to Public Comments on Amending<br />

General Permitting Regulations Relating to Habitat Conservation Plans, Safe Harbor<br />

Agreements and Candidate Conservation Agreements With Assurances. Final Rule.<br />

Federal Register: Volume 66, Number 14. Page 6483-6487.<br />

U. S. Fish and Wildlife Service (USFWS). 2001b. Biological/Conference Opinion Regarding<br />

the Effects of Operation of the Bureau of Reclamation’s Klamath Project on the<br />

Endangered Lost River Sucker (Deltistes luxatus), Endangered Shortnose Sucker<br />

(Chasmistes brevirostris), Threatened Bald Eagle (Haliaeetus leucocephalus) and<br />

Proposed Critical Habitat for the Lost River/Shortnose Suckers. U.S. Fish and Wildlife<br />

Service, Klamath Falls Fish and Wildlife Office, Klamath Falls, OR.<br />

U.S. Fish and Wildlife Service (USFWS). 2002. Biological/Conference Opinion Regarding the<br />

Effects of the Operation of the U.S. Bureau of Reclamation’s Proposed 10-year Operation<br />

Plan for the Klamath Project and its Effect on the Endangered Lost River Sucker<br />

(Deltistes luxatus), Endangered Shortnose Sucker (Chasmistes brevirostris), Threatened<br />

Bald Eagle (Haliaeetus leucocephalus) and Proposed Critical Habitat for the Lost<br />

River/Shortnose Suckers. U.S. Fish and Wildlife Service, Klamath Falls Fish and<br />

Wildlife Office, Klamath Falls, OR.<br />

U.S. Fish and Wildlife Service (FWS). 2003a. Finding for an Amended Petition To List the<br />

Westslope Cutthroat Trout as Threatened Throughout Its Range, 68 FR 46989 (August 7,<br />

2003).


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 47<br />

U. S. Fish and Wildlife Service (FWS). 2003b. Safe Harbor Agreements and Candidate<br />

Conservation Agreements With Assurances; Revisions to the Regulations. Federal<br />

Register: September 10, 2003 (Volume 68, Number 175); Proposed Rules. Page 53320-<br />

53327.<br />

U.S. Fish and Wildlife Service. 2003c. Endangered and Threatened Wildlife and Plants; Listing<br />

of the Central California Distinct Population Segment of the California<br />

Tiger Salamander; Reclassification of the Sonoma County and Santa Barbara County<br />

Distinct Populations From Endangered to Threatened; Special Rule. Federal Register:<br />

July 3, 2003 (Volume 68, Number 128); 50 CFR Part 17; RIN 1018-AI68; Page 39892-<br />

39893.<br />

U.S. Fish and Wildlife Service. 2003d. Endangered and Threatened Wildlife and Plants;<br />

Designation of Critical Habitat for the Gulf Sturgeon; Final Rule. Federal Register:<br />

March 19, 2003 (Volume 68, Number 53; Page 13369-13495; 50 CFR Part 226.<br />

U.S. Fish and Wildlife Service. 2004. Western sage grouse 69 Fed. Reg. 933, 935. Jan. 7, 2004.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1994.<br />

Endangered and threatened wildlife and plants: Notice of interagency cooperative policy<br />

for the ecosystem approach to the Endangered Species Act. Federal Register Doc. 94-<br />

16025, June 30, 1994.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1996a.<br />

Endangered species petition management guidance. Division of Endangered Species, U.<br />

S. Dept. of the Interior, Fish and Wildlife Service, Washington, D.C.; Endangered<br />

Species Division, National Marine Fisheries Service, National Oceanic and Atmospheric<br />

Administration, Dept. of Commerce, Silver Spring, MD.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1996b.<br />

cite for dps policy joint (61 FR 4722 (Feb. 7, 1996)


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 48<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1998.<br />

Habitat Conservation Plan assurances (“No Surprises”) rule. 50 CFR Part 17 and Part<br />

222: 8859, March 25, 1998.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1999a.<br />

Endangered and threatened wildlife and plants; Final listing priority guidance for fiscal<br />

year 2000. 64 Federal Register 57<strong>11</strong>4-57<strong>11</strong>9.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 1999b.<br />

Notice of availability of Final Endangered Species Consultation Handbook for<br />

procedures for conducting consultation and conference activities under Section 7 of the<br />

Endangered Species Act. 64 Federal Register 31285.<br />

U. S. Fish and Wildlife Service (FWS) and National Marine Fisheries Service (NMFS). 2003.<br />

Policy for evaluation of conservation efforts when making listing decisions. 50 CFR<br />

<strong>Chapter</strong> IV: 15100, April 28, 2003. Fish and Wildlife Service.<br />

General Accounting Office (GAO). 2002. Canada lynx survey. Unauthorized hair samples<br />

submitted for analysis. Statement of Ronald Malfi, Acting Managing Director, Office of<br />

Special Investigations. Testimony before the Committee on Resource, U.S. House of<br />

Representatives. GAO-02-496T.<br />

General Accounting Office (GAO) 2003. Endangered Species. Fish and Wildlife Service Uses<br />

Best Available <strong>Science</strong> to Make Listing Decisions, but Additional Guidance Needed for<br />

Critical Habitat Designations.<br />

Gerber, LR and GR VanBlaricom. 2001. Implications of three viability models for the<br />

conservation status of the western population of Stellar sea lions (Eumetopias jubatus).<br />

Biological Conservation 102: 261-269.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 49<br />

Gerber, LR and L. Hatch. 2002. Are we recovering? An evaluation of recovery criteria under<br />

the U.S. Endangered Species Act. Ecological Applications 12: 668-673.<br />

Halpern, E and J Wilson. 2003. A sellout, or just practical? He’s the biologist who pioneered<br />

use of land swaps to help builders develop acreage where endangered species live.<br />

Critics call him a traitor. Los Angeles Times, CA. March 14, 2003.<br />

Hanski, I. 1999. Metapopulation Ecology. Oxford University Press, Oxford, UK. 320 p<br />

Harris, FA. June 2, 2003. President’s Hook: Relevance in Public Policy. American Fisheries<br />

Society. http://www.fisheries.org/fisheries/F2806/President.htm<br />

Heppel, SS and LB Crowder. 1998. Prognostic evaluation of enhancement programs using<br />

population models and life history analysis. Bulletin of Marine <strong>Science</strong> 62: 495-507.<br />

Heppell, SS, LB Crowder, and DT Crouse. 1996. Models to evaluate headstarting as a<br />

management tool for long-lived turtles. Ecological Applications 6: 556-565.<br />

Hoekstra, J.M., W.F. Fagan and J. E. Bradley. 2002. A critical role for critical habitat in the<br />

recovery planning process? Not yet. Ecological Applications 12: 701-707.<br />

Holmes EE. 2001. Estimating risks in declining populations with poor data. Proc. Natl. Acad.<br />

Sci. USA 98:5072-77<br />

Holmes EE, Fagan WF. 2002. Validating population viability analyses for corrupted data sets.<br />

Ecology 83: 2379–2386.<br />

IUCN. International Union for the Conservation of Nature. 1994. IUCN Red list categories.<br />

21 p. Gland, Switzerland: IUCN Council.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 50<br />

Jenkins, CN, RD Powell, OL Bass Jr and SL Pimm. 2003. Demonstrating the destruction of the<br />

habitat of the Cape Sable seaside sparrow (Ammodramus maritimus mirabilis). Animal<br />

Conservation 6: 29-38.<br />

Kareiva PM, <strong>And</strong>elman S, Doak D, Elderd B, Groom M et al 1998. Using science in habitat<br />

conservation plans. www.nceas.ucsb.edu/projects/hcp.<br />

Kareiva PM, Marvier M, McClure M. 2000. Recovery and management options for<br />

spring/summer chinook salmon in the Columbia River Basin. <strong>Science</strong> 270: 977-79.<br />

Keith, DA, M. McCarthy, H. Regan, T. Regan, C. Bowles, C. Drill, C. Craig, B. Pellow, MA.<br />

Burgman, LL. Master, M. Ruckelshaus, B. Mackenzie, SJ. <strong>And</strong>elman, PR. Wade. 2004.<br />

Protocols for listing threatened species can forecast extinction. Ecology Letters, 7: <strong>11</strong>01-<br />

<strong>11</strong>08.<br />

Krahn, M.M., M. J. Ford, W. F. Perrin, P. R. Wade, R. P. Angliss, M. B. Hanson, B. L. Taylor,<br />

G. M. Ylitalo, M. E. Dahlheim, J. E. Stein, R. S. Waples. 2004. 2004 Status Review of<br />

Southern Resident Killer Whales (Orcinus orca) under the<br />

Endangered Species Act. Technical Memorandum NMFS-NWFSC-62. 73 p.<br />

Legislative History 1982. SENATE COMM. ON ENV’T AND PUB. WORKS, LEGISLATIVE<br />

HISTORY OF THE ENDANGERED SPECIES ACT OF 1973.<br />

Manson, C. 2003. Testimony of Craig Manson, Assistant Secretary for Fish and Wildlife and<br />

Parks, Department of the Interior, Before the subcommittee on Fisheries, Wildlife and<br />

Water of the Senate Committee on Environment and Public Works, regarding the


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 51<br />

designation of critical habitat under the Endangered Species Act. April 10, 2003,<br />

reprinted at http://laws.fws.gov/TESTIMON/2003/2003april10.html.<br />

Manson, C. 2004. TESTIMONY OF CRAIG MANSON, ASSISTANT SECRETARY FOR<br />

FISH AND WILDLIFE AND PARKS, DEPARTMENT OF THE INTERIOR, BEFORE<br />

THE HOUSE RESOURCES COMMITTEE, REGARDING H.R. 2933, THE CRITICAL<br />

HABITAT REFORM ACT OF 2003, April 28, 2004. reprinted at<br />

http://laws.fws.gov/TESTIMON/2004<br />

Mapes LV. 2001. Lynx-fur furor focuses on science role. Seattle Times, Dec. 30<br />

McCarthy, Michael, D. Keith, J. Tietjen, M. A. Burgman, M. Maunder, L. Master, B. Brooks, G.<br />

Mace, H. Possingham, R. Medellin, S. <strong>And</strong>elman, H. Regan, T. Regan, M. Ruckelshaus.<br />

2004. Comparing predictions of extinction risk using models and subjective judgement.<br />

Acta Oecologica, 26: 67-74.<br />

McCarty, JP. 2001. Ecological consequences of recent climate change. Conservation Biology<br />

15: 320-331.<br />

McClure MM, Holmes EE, Sanderson BL, Jordan CE. 2002. A large-scale, multi-species<br />

assessment: Anadromous salmonids in the Columbia River Basin. Ecol. Appl. In press<br />

McElhany P, Ruckelshaus MH, Ford MJ, Wainwright T, Bjorkstedt E. 2000. Viable salmonid<br />

populations and the recovery of evolutionarily significant units. U.S. Dept. Commerce,<br />

NOAA Tech. Memorandum NMFS-NWFSC-42, 156 pp<br />

Miami Herald, The. 2003. Florida panther fatalities rising on state’s roadways. Miami, FL.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 52<br />

June 18, 2001.<br />

Miller, G. 1996. Ecosystem management: Improving the Endangered Species Act. Ecological<br />

Applications 6: 715-717.<br />

Mills LS. 2002. False samples are not the same as blind controls. Nature 415: 471<br />

Morris W, Doak D, Groom M, Kareiva PM, Fieberg J et al. 1999. A Practical Handbook for<br />

Population Viability Analysis. Washington, D.C.: The Nature Conservancy<br />

Myer 2001. Community Politics and Endangered Species Protection, in Protecting J. Shogren<br />

and Tschirhart (eds.) ENDANGERED SPECIES IN THE UNITED STATES,<br />

Cambridge Univ. Press, London.<br />

Myers, RA, Simon A. Levin, Russell Lande, Frances C. James, William W. Murdoch, Robert T.<br />

.<br />

Paine 2004. Hatcheries and endangered salmon. <strong>Science</strong> 303:1980.<br />

National Marine Fisheries Service (NMFS). 1991. Notice of policy: Policy on applying the<br />

definition of species under the Endangered Species Act to Pacific salmon. Federal<br />

Register [Docket 910248-1255, 20 November 1991] 56(224): 58612–-58618.<br />

National Marine Fisheries Service (NMFS). 1992. Recovery planning guidelines. National<br />

Oceanic and Atmospheric Administration, National Marine Fisheries Service, Office of<br />

Protected Resources.<br />

National Marine Fisheries Service (NMFS). 1993. Interim policy on artificial propagation of<br />

Pacific salmon under the Endangered Species Act. Federal Register [Docket 92<strong>11</strong>86-<br />

2286, 5 April 1993] 58(63): 17573–17576.<br />

National Marine Fisheries Service (NMFS) 1993d. 58 FR 3121 (Jan. 7, 1993)


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 53<br />

National Marine Fisheries Service (NMFS) 1998. NMFS’ Biological Opinion on Berg Brothers,<br />

Water Withdrawal Columbia River, 1998, reprinted<br />

athttp://www.nwr.noaa.gov/1publcat/bo/1998/19970<strong>11</strong>89_berg_irrigation_12-01-<br />

1998.pdf<br />

National Marine Fisheries Service (NMFS). 1999. Biological Opinion on Columbia River<br />

Federal Navigation Channel Deepening 1999, reprinted at<br />

http://www.nwr.noaa.gov/1publcat/bo/1999/osb1999-0270.<br />

National Marine Fisheries Service (NMFS) 2000a. Reinitiation of Consultation on the Federal<br />

Columbia River Power System (FCRPS) Including the Juvenile Fish Transportation<br />

Program and 19 Bureau of Reclamation Projects in the Columbia Basin, 2000, reprinted<br />

at http://www.nwr.noaa.gov/1publcat/bo/2000/2000.html).<br />

National Marine Fisheries Service (NMFS) 2000b. Basinwide Salmon Recovery Strategy<br />

December 2000, reprinted at http://www.salmonrecovery.gov/strategy.shtml.<br />

National Marine Fisheries Service (NMFS) 2000c. Biological Opinion. Reinitiation of<br />

Consultation on Operation of the Federal Columbia River Power System,<br />

Including the Juvenile Fish Transportation Program, and 19 Bureau of Reclamation<br />

Projects in the Columbia Basin. Available at:<br />

http://www.nwr.noaa.gov/1hydrop/hydroweb/docs/Final/2000Biop.html<br />

National Marine Fisheries Service (NMFS) 2001. Biological Opinion. Ongoing Klamath<br />

Project Operations. National Marine Fisheries Service, Southwest Region, National<br />

Oceanic and Atmospheric Association, Long Beach, CA. April 6, 2001.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 54<br />

National Marine Fisheries Service (NMFS) 2002a. Biological Opinion on Columbia River<br />

Federal Navigation Channel Deepening 2002, reprinted in<br />

http://www.nwr.noaa.gov/1publcat/bo/2002/200101414_federal_navigation_05-20-<br />

2002.pdf<br />

National Marine Fisheries Service (NMFS) 2002b. 12-Month Finding Petition to List Southern<br />

Resident Killer Whales Not Warranted (67 FR 4413 (July 1, 2002).<br />

National Marine Fisheries Service (NMFS). 2002a. Endangered and threatened species: Findings<br />

on petitions to delist Pacific salmonid ESUs. Federal Register [Docket No. 020205024-<br />

2024-01] 67(28): 6215–6220.<br />

National Marine Fisheries Service (NMFS). 2002d. Biological Opinion, Klamath Project<br />

Operations. National Marine Fisheries Service, Southwest Region, National Oceanic and<br />

Atmospheric Association, Long Beach, CA. May 31, 2002.<br />

National Marine Fisheries Service (NMFS). 2003. Updated Status of Federally Listed ESUs of<br />

West Coast Salmon and Steelhead. West Coast Salmon Biological Review Team,<br />

Northwest Fisheries <strong>Science</strong> Center, Seattle, WA 98<strong>11</strong>2, and Southwest Fisheries<br />

<strong>Science</strong> Center, Santa Cruz, CA 95060. July 2003.<br />

National Marine Fisheries Service (NMFS). 2004a. Endangered and threatened wildlife and<br />

plants: Proposed threatened status for southern resident killer whales. 50 CFR Part 223;<br />

December 22, 2004.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 55<br />

National Marine Fisheries Service (NMFS). 2004b. Proposed policy for the consideration of<br />

hatchery-origin fish in ESA listing determinations. 69 FR 31354; June 3, 2004.<br />

National Marine Fisheries Service (NMFS). 2005. Policy on the consideration of hatchery-<br />

origin fish in Endangered Species Act listing determinations for Pacific salmon and<br />

steelhead. 50 CFR Parts 223 and 224. Federal Register Vol. 70 No. 123. Tuesday, June<br />

28, 2005.<br />

National Research Council (NRC) ed. 1995. <strong>Science</strong> and the Endangered Species Act.<br />

Washington, DC: National Academy Press<br />

National Research Council (NRC) ed. 2002. Scientific evaluation of Biological Opinions on<br />

endangered and threatened fishes in the Klamath River Basin. Interim Report from the<br />

Committee on Endangered and Threatened Fishes in the Klamath River Basin. National<br />

Academy Press, Washington, D.C.<br />

National Research Council (NRC) ed. 2003. Water Resources Management, Instream Flows,<br />

and Salmon Survival in the Columbia River Basin, Washington, Scope of Work and<br />

Budget.<br />

National Research Council (NRC) ed. 2004a. Managing the Columbia River: Instream Flows,<br />

Water Withdrawals, and Salmon Survival, March 31, 2004. The National Academy of<br />

<strong>Science</strong>s, National Academy Press, Washington, D.C.<br />

National Research Council (NRC) ed. 2004b. Endangered and Threatened Fishes in the<br />

Klamath River Basin: Causes of Decline and Strategies for Recovery. The National<br />

Academy of <strong>Science</strong>s, National Academy Press, Washington, D.C.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 56<br />

NatureServe. 2003. http://www.natureserve.org/<br />

Nowack, RM. 1992. The red wolf is not a hybrid. Conservation Biology 6: 593-595.<br />

Office of Management and Budget (OMB). 2003. Circular A-4, September 17, 2003<br />

Patlis 2001. Patlis, Jason, Paying Tribute to Joseph Heller with the Endangered Species Act:<br />

When Critical Habitat Isn’t, 20 Stan. Envtl L.J. 133 (2001)<br />

Paulsen, CM and RA Hinrichsen. 2002. Experimental management for Snake River spring-<br />

summer chinook (Oncorhynchus tshawytscha): trade-offs between conservation and<br />

learning for a threatened species. Canadian Journal of Fisheries and Aquatic <strong>Science</strong>s 59:<br />

717-725.<br />

Peterson G, Cumming G, Carpenter S. 2003. Scenario planning: a tool for conservation in an<br />

uncertain world. Conservation Biology 17: 358-366.<br />

Pfeifer, S. 2003. New plan endangers the same projects; latest proposal to protect the<br />

gnatcatcher could halt a toll road extension and the Rancho Mission Viejo home<br />

development. Los Angeles Times, CA. April 25, 2003.<br />

Pianin, E. 2003. Judge orders river level lowered—Endangered Species Act takes precedence,<br />

ruling says. Washington Post, Washington, D.C. August 7, 2003.<br />

Puget Sound Technical Recovery Team (PSTRT). 2002. Planning ranges and preliminary<br />

guidelines for the delisting and recovery of the Puget Sound Chinook salmon<br />

Evolutionarily Significant Unit. April 30, 2002. (available at:<br />

http://www.nwfsc.noaa.gov/trt/trtpopESU.pdf)


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 57<br />

Pyare, S. and J. Berger. 2003. Beyond demography and delisting: ecological recovery for<br />

Yellowstone’s grizzly bears and wolves. Biological Conservation <strong>11</strong>3: 63-73.<br />

Ralls, K., Demaster, D. P. & Estes, J. A. 1996 Developing a criterion for delisting the southern<br />

sea otter under the U.S. Endangered Species Act. Conservation Biology 10, 1528-1537.<br />

Rinkevich, SE and SC Leon. 2000. Stakeholders assist species recovery in the southwest.<br />

Endangered Species Bulletin 25: 28-29.<br />

Roby, DD, DE Lyons, DP Craig, K Collis and GH Visser. 2003. Quantifying the effect of<br />

predators on endangered species using a bioenergetics approach: Caspian terns and<br />

juvenile salmonids in the Columbia River estuary. Canadian Journal of Zoology 81: 250-<br />

265.<br />

Rohlf, D.J. 1989. The Endangered Species Act: A Guide to Its Protections and <strong>Implementation</strong>,<br />

Stanford Environmental Law Society, Stanford, CA.<br />

Rohlf, D.J. 2001. Jeopardy Under the Endangered Species Act: Playing A Game Protected<br />

Species Can’t Win, 41 Washburn L. Rev. <strong>11</strong>4, 158 (2001).<br />

Rosenberg, AA. 2002. The precautionary approach from a manager’s perspective. Bull. Marine<br />

Sci. 70: 577-588.<br />

Rubidge, E, P Corbett and EB Taylor. 2001. A molecular analysis of hybridization between<br />

native westslope cutthroat trout (Oncorhynchus clarki lewisi) and introduced rainbow<br />

trout (O. mykiss) in southeastern British Columbia. Journal of Fish Biology 59: 42-54.<br />

Rubidge, E. 2003. A genetic analysis of hybridization between native westslope cutthroat trout<br />

(Oncorhynchus clarki lewisi) and introduced rainbow trout (O. mykiss) in the upper


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 58<br />

Kootenay River drainage of southeastern British Columbia. MSc Thesis, University of<br />

British Columbia.<br />

Ruckelshaus, M., P. Levin. J. Johnson and P. Kareiva. 2002a. The Pacific salmon wars: What<br />

science brings to the challenge of recovering species. Annual Review of Ecology and<br />

Systematics. 33:665-706.<br />

Ruckelshaus, M., P. McElhany and M. J. Ford. 2002b. Recovering species of conservation<br />

concern: are populations expendable? Pp. 305-329. In: P. Kareiva and S. Levin (eds.)<br />

The Importance of Species: Perspectives on expendability and triage. Princeton<br />

University Press, Princeton, NJ.<br />

Ruckelshaus, M., P. McElhany, M. McClure and S. Heppell. 2004. Chinook salmon in<br />

Puget Sound: Effects of spatially correlated catastrophes on persistence. Pp. 208-<br />

218, In R. Ackakaya, M. Burgman, O. Kindvall, C.C. Wood, P. Sjogren-Gulve, J. S.<br />

Hatfield and M. A. McCarthy (eds.) Species Conservation and Management: Case<br />

Studies. Oxford Univ. Press.<br />

Sacramento Bee. 2003a. South county adrift. Supervisor’s plan: To hell with the hawk.<br />

Sacramento, CA. August 10, 2003.<br />

Sacramento Bee. 2003b. Not so critical. Habitat designation has become distraction.<br />

Sacramento, CA. August 22, 2003.<br />

Schwartz, MW. 1999. Choosing the appropriate scale of reserves for conservation. Annual<br />

Review of Ecology and Systematics 30: 83-108.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 59<br />

Seattle Times. 2002. Leery of grizzly tests, legislator demands review of all endangered-species<br />

studies, Jan. 10<br />

Shelden, KEW, DP DeMaster, DJ Rugh, and AM Olson. 2001. Developing classification<br />

criteria under the U.S. Endangered Species Act: Bowhead whales as a case study.<br />

Conservation Biology 15: 1300-1307.<br />

Sinden 2004. THE ECONOMICS OF ENDANGERED SPECIES: WHY LESS IS MORE IN<br />

THE ECONOMIC ANALYSIS OF CRITICAL HABITAT DESIGNATIONS, 28 Harv.<br />

Envtl. L. Rev. 129 2004.<br />

Springer, AM, J. A. Estes, G. B. van Vliet, T. M. Williams, D. F. Doak, E. M. Danner, K. A.<br />

Forney, and B. Pfister. 2003. Sequential megafaunal collapse in the North Pacific<br />

Ocean: An ongoing legacy of industrial whaling? Proceedings of the National Academy<br />

of <strong>Science</strong>s 100: 12223-12228.<br />

Stokstad E. 2002. Fur flies over charges of misconduct. <strong>Science</strong> 295:250-51<br />

Strassel KA. 2002. The missing lynx. The Wall Street Journal, Jan. 24<br />

Taylor, EB, MD Stamford, and JS Baxter. 2003. Population subdivision in westslope cutthroat<br />

trout (Oncorhynchus clarki lewisi) at the northern periphery of its range: evolutionary<br />

inferences and conservation implications. Molecular Ecology 12: 2609-2622.<br />

Taylor, M., K. Suckling and J. Rachlinski. 2003. Critical habitat significantly enhances<br />

endangered species recovery. Center for Biological Diversity, Tucson, AZ. Report<br />

available at: www.biologicaldiversity.org/swcbd/programs/policy


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 60<br />

Tear TH, Scott JM, Hayward PH, Griffith B. 1993. Status and prospects for success of the<br />

Endangered Species Act: A look at recovery plans. <strong>Science</strong> 262:976-97<br />

Verhovek SH. 2002. “Saving” wild salmon’s bucket-born cousins. New York Times, Feb. 4<br />

Wainwright, T. C. and R. G. Kope. 1999. Methods of extinction risk assessment developed for<br />

U.S. West Coast salmon. ICES J. Mar. Sci. 56:444-448.<br />

Waples RS. 1991. Pacific salmon (Oncorhynchus spp.) and the definition of “species” under the<br />

Endangered Species Act. Mar. Fish. Rev. 53:<strong>11</strong>-22<br />

Waples RS. 1995. Evolutionarily significant units and the conservation of biological diversity<br />

under the Endangered Species Act. In Evolution and the aquatic ecosystem: defining<br />

unique units in population conservation, ed. JL Nielsen, pp. 8-27. American Fisheries<br />

Society Symposium, 17th, Bethesda, MD: American Fisheries Society<br />

Waples RS. 1999. Dispelling some myths about hatcheries. Fisheries 24:12-21<br />

Waples, R.S., Teel, D.J., Myers, J.M., Marshall, A.R., 2004. Life-history divergence in Chinook<br />

salmon: Historic contingency and parallel evolution. Evolution 58, 386-403.<br />

Washington Post. 2004. June 21.<br />

Wayne, RK et al. 1992. Mitochondrial DNA variability of the gray wolf: genetic consequences<br />

of population and habitat fragmentation. Conservation Biology 6: 559-569.<br />

Wayne, RK and SM Jenks. 1991. Mitochondrial DNA analysis implying extensive<br />

hybridization of the endangered red wolf, Canis rufus. Nature 351: 565-568.<br />

Weber B. 2002. Wildlife society clarifies its role in lynx studies. The Wall Street Journal, Feb.<br />

12.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 61<br />

Willamette-Lower Columbia Technical Recovery Team (WLCTRT). 2003. Interim report on<br />

viability criteria for Willamette and Lower Columbia Basin Pacific salmonids. (available<br />

at: http://www.nwfsc.noaa.gov/trt/viability_report.htm)<br />

Wilson, J. 2003a. Riverside County set to approve big land plan, a species trade-off; with<br />

threats of lawsuits from both sides of the preservation debate, the Board of Supervisors is<br />

expected to act today on 1.2 million acres. Los Angeles Times, CA. June 17, 2003.<br />

Wilson, J. 2003b. U.S. Officials cut protected habitat throughout the state. Fish and Wildlife<br />

Service cites economic concerns in plan that eliminates preserves for endangered species<br />

in six counties. Appeal vowed. Los Angeles Times, CA. August 7, 2003.<br />

Wilson, PJ, S Grewal, ID Lawford, JNM Heal, AG Granacki, D Pennock, JB Theberge, MT<br />

Voight, W Waddell, RE Chambers, PC Paquet, G Goulet, D Cluff, BN White. 2000.<br />

DNA profiles of the Canadian wolf and the red wolf provide evidence for a common<br />

evolutionary history independent of the gray wolf. Canadian Journal of Zoology 78:<br />

2156-2166.<br />

Wondolleck JM, Yaffee SL. 2000. Making Collaboration Work. Lessons from Natural Resource<br />

Management. Washington, DC: Island Press<br />

Zink, RM, GF Barrowclough, JL Atwood, and RC Blackwell-Rago. 2000. Genetics, taxonomy,<br />

and conservation of the threatened California Gnatcatcher. Conservation Biology 14:<br />

1394-1405.


Number of pertinent articles<br />

Search terms<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 62<br />

Figure <strong>11</strong>.1 The number of articles published since 1980 that address science questions posed in<br />

the Endangered Species Act. The published articles are those that occur in the ISI Web of<br />

<strong>Science</strong> <strong>Science</strong> Citation Index Expanded database between 1980 and September, 2003<br />

(Thompson ISI, 2003). Search terms pertaining to each science question are listed below each<br />

panel; articles that emerged for more than one search term within a question category were only<br />

counted once.<br />

1980-1983<br />

1984-1987<br />

1988-1991<br />

1.<br />

What is a<br />

“listable” species?<br />

1992-1995<br />

1996-1999<br />

2000-2003<br />

Distinct population segment*<br />

Endangered Species Act<br />

Evolutionarily significant unit<br />

ESUs<br />

1980-1983<br />

1984-1987<br />

1988-1991<br />

2.<br />

What is its<br />

risk of extinction?<br />

1992-1995<br />

1996-1999<br />

2000-2003<br />

Population viability analysis<br />

Population viability analyses<br />

Viability analys*<br />

Risk of extinction<br />

Extinction risk<br />

1980-1983<br />

1984-1987<br />

1988-1991<br />

3.<br />

What is critical habitat<br />

for the listed species?<br />

1992-1995<br />

Critical habitat*<br />

Essential habitat*<br />

1996-1999<br />

2000-2003<br />

1980-1983<br />

1984-1987<br />

1988-1991<br />

4.<br />

Does the action “jeopardize”<br />

the species’ existence?<br />

1992-1995<br />

1996-1999<br />

2000-2003<br />

Jeopard* AND endangered species<br />

Section 7 AND endangered


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 63<br />

Table <strong>11</strong>.1. Key science-related provisions within the Endangered Species Act that would benefit<br />

from amplification<br />

ESA provision <strong>Science</strong>-related<br />

question addressed by<br />

provision<br />

Further work needed on<br />

analyses pertaining to<br />

provision<br />

4(a) Listing Is there a “species”? Improved definition of<br />

4(a) Listing What is the species’<br />

4(b) Critical habitat<br />

designation<br />

risk of extinction?<br />

What habitat features<br />

are essential to species’<br />

conservation and how<br />

much habitat is needed<br />

the “distinct population<br />

segment/evolutionarily<br />

significant unit” concept<br />

Improved definition of<br />

time scales, attention to<br />

multiple indicators of<br />

risk, methods of<br />

estimating rates of<br />

reproduction of at-risk<br />

species<br />

Relationship between<br />

habitat quality/quantity<br />

and species extinction<br />

risk<br />

Further work needed<br />

on application of<br />

provision<br />

Agency guidance on<br />

how to address<br />

hybridization, definition<br />

of taxonomic<br />

“significance,” and<br />

artificially propagated<br />

individuals<br />

Agency guidance on<br />

consideration of<br />

extinction risk and<br />

“significant portion” of<br />

range<br />

Agency guidance on<br />

designating critical<br />

habitat, how to weigh<br />

benefits/costs; consider


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 64<br />

7(a)(2) Federal<br />

consultation<br />

10(a)(1)(B) Habitat<br />

Conservation Plans<br />

for conservation? sequence of application<br />

What effect will a<br />

particular action have<br />

on species’ survival or<br />

recovery?<br />

Does an action result in<br />

take, and if so, how<br />

much?*<br />

4(f) Recovery planning What are the<br />

characteristics of a<br />

recovered species?<br />

What factors are<br />

limiting recovery?<br />

What habitat is<br />

essential to recovery?<br />

Relative importance of<br />

different limiting factors<br />

in extinction risk; how<br />

effects of individual<br />

actions relate to whole<br />

population/species<br />

impacts<br />

All of the above<br />

More-open science<br />

process in section 7<br />

consultations; guidance<br />

on considering<br />

piecemeal vs. whole life-<br />

cycle approach<br />

More public<br />

participation in policy<br />

oversight of the planning<br />

process<br />

* This question also arises in section 9 enforcement actions, which are not addressed in this chapter.


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 65<br />

Box <strong>11</strong>.1. Examples of the role of science in implementation of section 7 limitations on federal<br />

actions under the Endangered Species Act<br />

Dealing with scientific uncertainty. The manner in which inevitable scientific uncertainty<br />

is incorporated into section 7 consultations is key to using science to inform sound<br />

decisions. In the high-profile case of Bureau of Reclamation operations of the Klamath<br />

River Basin water management project, the National Research Council was brought in to<br />

help resolve what many characterized as a scientific dispute (Cooperman and Markle<br />

2003). The Council reviewed 3 analyses: (1) the Bureau’s proposed action for operating<br />

the project and its predicted effects on listed sucker fish and coho salmon, (2) NMFS’ and<br />

USFWS’ 2001 biological opinions that found that the proposal would jeopardize the<br />

species and offered reasonable and prudent alternatives that would avoid jeopardy (NMFS<br />

2001, USFWS 2001b), and (3) the Bureau’s and the Services’ 2002 revisions to their<br />

assessments and biological opinions on a new set of proposed actions revised based on a<br />

court order (NMFS 2002d, USFWS 2002). The NRC’s final report (NRC 2004) highlights<br />

several recommendations aimed at reducing the uncertainty in the biological conclusions<br />

by the Services that the Bureau’s proposed actions will not jeopardize the listed species.<br />

Such recommendations include (1) the USFWS and NMFS are urged to complete recovery<br />

plans for the 2 species that should identify how research and monitoring will support<br />

species recovery and facilitate identification of what actions are allowed under section 7<br />

and 10 consultations, (2) scientists should be allowed sufficient time to publish key<br />

research findings in peer-reviewed scientific journals, (3) a diverse team of “cooperators”<br />

should be convened for designing ecosystem-based management actions that have local


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 66<br />

support for implementation, and (4) experiments should be conducted to test the<br />

effectiveness and feasibility of specific remediation strategies (NRC 2002 and 2004).<br />

In another example, NMFS was thwarted in an attempt to deal with uncertainties<br />

associated with future allocation of necessary conservation actions among sectors in the<br />

Columbia River Basin. The Federal Columbia River Power System is a system of 13 dams<br />

and reservoirs on the Columbia and Snake Rivers in Oregon, Washington, and Idaho. The<br />

basin is home to 12 species of endangered salmon and steelhead whose migrations are<br />

affected by operation of the power system. In 2000, NMFS issued a biological opinion on<br />

operation of the power system, recognizing that a recovery plan was not yet in place and<br />

that any recovery of Columbia River salmon and steelhead would require contributions<br />

from many sectors (NMFS 2000a). In an effort to be clear about how it was allocating the<br />

conservation burden, NMFS and the other federal agencies involved presented a<br />

conceptual recovery plan (NMFS 2000b). That plan described, for example, the<br />

assumptions that would have to be made about continued harvest restrictions into the<br />

future if the sum of impacts on listed fish was to avoid jeopardy. This opinion was<br />

invalidated by a district court finding that the agency improperly relied on assumed future<br />

actions that were not “reasonably certain to occur” (National Wildlife Federation v. NMFS,<br />

CR 01-640-RE [OR 2003]), leaving in question the ability of the Services to consider the<br />

“big picture” when section 7 biological opinions have implications for allocation of take.<br />

Considering actions in isolation. For many species it is the cumulative effect of many<br />

actions that have led to their imperilment and it is difficult for the Services in a section 7<br />

consultation to make the case that a single small action, when added to the many other


Chap. <strong>11</strong>, <strong>Science</strong> and <strong>Implementation</strong>, (ESA Vol. 2) Ruckelshaus and Darm, 67<br />

small actions, jeopardizes the species’ continued existence. In the case of water<br />

withdrawals from the Columbia River, NMFS did issue a jeopardy opinion to the Corps of<br />

Engineers on the basis that Columbia River flows were already below species’ needs in<br />

many years, the cumulative impact of withdrawals contributed to those low flows, and<br />

there was no mechanism in place to limit future withdrawals. Even though the withdrawal<br />

under consideration was very small compared to overall flows in the Columbia, NMFS<br />

concluded the proposed action would jeopardize Columbia River salmon and steelhead<br />

because of the cumulative effect of past and future withdrawals (NMFS 1998). This<br />

decision stirred considerable controversy in the Basin, leading Washington’s Department<br />

of Ecology to appeal to the NRC, asking the NRC to review the science supporting flow<br />

levels in the Columbia. Although the question put to the panel was framed in terms of the<br />

incremental risk posed by a very small incremental degradation in flows, the panel resisted<br />

being drawn into answering the narrow question. In its preliminary findings, the panel<br />

appears to support the analysis that because flows currently are inadequate, even small<br />

increases in water withdrawals will increase risk (NRC 2004).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!