Show new navigation
On
Columbia Basin Fish and Wildlife Program Columbia Basin Fish and Wildlife Program
Columbia Basin Fish and Wildlife Program Columbia Basin Fish and Wildlife Program
RSS Feed for updates to Proposal RMECAT-2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River Follow this via RSS feed. Help setting up RSS feeds?

Proposal Summary

Proposal RMECAT-2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River

View the dynamic Proposal Summary

This Proposal Summary page updates dynamically to always display the latest data from the associated project and contracts. This means changes, like updating the Project Lead or other contacts, will be immediately reflected here.

Download a snapshot PDF

To view a point-in-time PDF snapshot of this page, select one of the Download links in the Proposal History section. These PDFs are created automatically by important events like submitting your proposal or responding to the ISRP. You can also create one at any time by using the PDF button, located next to the Expand All and Collapse All buttons.


Archive Date Time Type From To By
Download 7/30/2010 5:12 PM Status Draft ISRP - Pending First Review <System>
10/15/2010 5:57 PM Status ISRP - Pending First Review ISRP - Pending Final Review <System>
1/19/2011 2:47 PM Status ISRP - Pending Final Review Pending Council Recommendation <System>
7/8/2011 1:28 PM Status Pending Council Recommendation Pending BPA Response <System>

This online form is dynamically updated with the most recent information. To view the content as reviewed by the ISRP and Council for this review cycle, download an archived PDF version using the Download link(s) above.

Proposal Number:
  RMECAT-2003-054-00
Proposal Status:
Pending BPA Response
Proposal Version:
Proposal Version 1
Review:
RME / AP Category Review
Portfolio:
RM&E Cat. Review - Artificial Production
Type:
Existing Project: 2003-054-00
Primary Contact:
Michael Blouin
Created:
6/7/2010 by (Not yet saved)
Proponent Organizations:
Oregon State University

Project Title:
Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
 
Proposal Short Description:
TWO MAIN OBJECTIVES
I: Estimate the fitness effects of raising steelhead in a hatchery, and the effects of those fish on wild populations of steelhead in the Hood River.
II: Identify mechanisms causing hatchery fish to become different from wild fish. Suggest ways to alleviate the problem.

IMPORTANCE: Accurate estimates of fitness are necessary for modeling demographic effects of hatcheries on wild populations. Understanding mechanisms causing decline may identify solutions.
 
Proposal Executive Summary:
BACKGROUND
The Hood River supports two populations of steelhead, a summer run and a winter run. They spawn only above the Powerdale Dam, which was a complete barrier to all salmonids. Steelhead samples have been collected at the dam every year by the ODFW. Since 1991 every adult passed above the dam has been measured, and sampled. Therefore, we have age data (from scales) and a DNA sample (from scales or fin snips) from every adult steelhead that went over the dam to potentially spawn in the Hood River from 1991 to the present. The dam was removed in June 2010, so including this year’s samples, we will eventually have a 19-year pedigree when it is complete. Similar numbers of hatchery (H) and wild (W) fish were passed above the dam during the last decade. During the 1990's "old" domesticated hatchery stocks of each run (multiple generations in the hatchery, out-of-basin origin) were phased out, and conservation hatchery programs were started for the purpose of supplementing the two wild populations (hereafter "new" hatchery stocks, “Hnew”; often referred to as “F1s”). In the following I use “wild” or “wild-born” to mean born in the wild, regardless of parentage (same as “natural origin” or “NOR” of some authors). The winter-run Hnew were started in 1991, and the summer-run Hnew were started in 1997. These samples give us the ability to estimate, via microsatellite-based pedigree analysis, the lifetime reproductive success (adult to adult production) of hatchery (H) and wild (W) fish for two populations (summer and winter), over multiple brood years (e.g. see Figs. 1-3 in “Accomplishments” section). We now have a three-generation pedigree that is complete for all anadromous winter-run fish, and we will eventually complete it to four generations. Note, however, that we are missing many parents, usually fathers. We have recently shown that few of these missing parents could be residualized hatchery fish, so we conclude they are mostly wild residents (Christie et al., in review). Although the dam has been removed, ODFW staff will continue to sample from fish captured at weirs in the basin. We will archive those samples for future use.

We have been using the Hood River system to work on the following questions:

(1) Do old-stock and/or first-generation (Hnew) hatchery fish differ from wild fish in relative reproductive success (RRS)?

(2) If there is a difference, is it genetically based?

(3) Given 2 is true, are those hatchery fish dragging down the fitness of the wild population?

(4) What are the mechanisms causing hatchery fish to be genetically different from wild fish?

We previously showed: (1) old, domesticated hatchery fish in the Hood River have very low fitness, with RRS = 10-30% that of wild fish. (2) Hnew, winter-run hatchery fish in the Hood River have about 85% the fitness of wild fish. (2) second-generation hatchery fish have about 55% the fitness of first-generation fish when all were raised side-by-side in the same hatchery. Thus there appears to be a very strong genetic effect of hatchery culture. (3) The RRS of wild-born fish depends on whether their parents were wild or hatchery fish, with offspring of naturally-occurring HxH crosses having <40% the RRS of offspring of WxW crosses. These wild fish presumably all experienced the same environment, so the difference is probably genetic as well. Furthermore, this latter result suggests that even a full generation of natural selection in the wild did not erase the hatchery effect, which in turn implies that these first-generation fish could be dragging down the fitness of the wild population. We also studied related topics such as hatchery effects on the effective population size, reproduction by residualized hatchery fish, and tests of some hypothesized mechanisms that could cause the hatchery fitness decline. See details in “Accomplishments” section.

PRIMARY GOALS OF CONTINUING WORK AND WHY IS IMPORTANT

Objective I: fulfill our original mission of estimating the fitness effects of raising fish in a hatchery, and of the effects of those fish on wild populations in the Hood River.

Project deliverables under objective I
Complete the full 19-year pedigree for summer-run and winter-run steelhead in the Hood River (1991-2010; four generations). Analyze the effects of varying degrees of hatchery background on fitness of fish in the wild and when used as broodstock in the hatchery. Here we have a complex set of genetic backgrounds because fish of differing ancestry have been spawning in the wild and also being used as broodstock. We propose teasing apart those various effects to understand to what extent the hatchery program has impacted the wild population. For the winter run, the players in the system are wild fish, resident fish, and first-generation hatchery fish. For the summer run, the players are wild fish, resident fish, an old domesticated hatchery stock (Skamania) from the early 1990s, and a first-generation stock from the late 1990s.

Importance
We need point estimates of hatchery/wild RRS for multiple run years in order understand variability of estimates. We also need studies whose design allows us to distinguish between genetic effects and those that are purely an environmental phenotypic effect of hatchery culture. Such data are essential for properly parameterizing models that predict the long-term demographics of supplemented populations. Testing the effects of bringing fish of different ancestry into the hatchery as broodstock is important because in most supplementation programs one cannot know the ancestry of wild-born fish taken for broodstock.


Objective II: identify mechanisms that cause hatchery fish to become different from wild fish, and suggest ways to alleviate the problem.

We recognize that this is a very large objective. Our initial foray into this question will involve two specific questions (deliverables). In the first, we propose searching for genes that have changed as a result of directional selection by the hatchery. In the second, we study one particular genetic system (immunity genes of the major histocompatibility complex, MHC) in which relaxed selection on mate choice could have resulted in lower fitness of hatchery fish.

Project deliverables under objective II
1. Analysis of gene expression in identically-raised juveniles of H and W background: includes identification of genes that are differentially expressed between types of fish and an analysis of the types of traits that might have been under selection.

2. (a) Test whether MHC genotype has a detectable effect on fitness of H and W steelhead in the Hood River.
(b) Estimate the improvement in performance of hatchery fish that would be achieved by choosing broodstock based on MHC genotype.

Importance
It is now clear that low fitness of hatchery fish is a general phenomenon. So what do we do about it? Keeping hatchery and wild fish separate will be difficult in most river systems. An alternate solution would be to figure out how to make hatchery fish that don’t impose a genetic load on wild populations. If we understood the mechanisms causing genetic change in hatcheries, then it might be possible to modify hatchery practices to produce fish that are compatible with wild populations. Thus, identifying those mechanisms is a critical area of research.

HOW THE WORK WILL BE ACCOMPLISHED
Objective I.
We will genotype the remaining few thousand fish samples that have not yet been done, plus those that need to be re-done to fill in gaps. Then we will attempt a global analysis of fitness in the wild and in the hatchery as a function of different amounts of hatchery background. The lab methods are the same as we have used all along. We may need to develop some new statistical methods for the global analysis, but parentage and kinship methods are a particular area of interest in my lab (e.g. Blouin et al., 1995; Blouin, 2003; Araki and Blouin, 2005; Christie et al., 2010; Christie et al., in review).

Objective II.
The gene expression work will be done using the Illumina 1G, next generation sequencer at OSU, in collaboration with my colleague at OSU, Todd Mockler (details under “Problem Statement”). We have already raised and frozen the fish to be analyzed.

The MHC work involves standard PCR and sequencing methods (details in “Problem Statement”). We already have in hand DNA samples from fish whose fitness we have estimated. Therefore, the proposed hypothesis tests are easily accomplished. Tie-in projects with disease-challenges and parasite surveys will be handled by colleagues who specialize in those areas (Jeri Bartholomew, OSU Microbiology, and Kym Jacobson, NOAA Fisheries) (more details in “Problem Statement”). My lab will sequence the same fish at MHC.

WHERE THE WORK WILL BE DONE, FOR HOW LONG, AND BY WHOM
We propose four years of work (FY 2012-2015). The genetics will be carried out by Michael Blouin’s lab at Oregon State University. Our lab includes a technician, a postdoc (currently Dr. Mark Christie) and a graduate student. This project is coordinated with the Hood River steelhead hatchery and research program, administered and implemented by the Oregon Department of Fish and Wildlife and the Warm Springs Tribes. We also collaborate with staff at Parkdale and Oak Springs hatchery to do experimental crosses and raise fish. We collaborate with the laboratories of Drs. Todd Mockler, J. Bartholomew and K. Jacobson on the gene expression and fish disease work.

HOW EFFECTIVENESS OF YOUR WORK WILL BE MONITORED
For each deliverable, the benchmark to evaluate results is the production of data that estimate the parameter of interest and/or test the hypothesis stated. Our standard of data quality is publication in peer-reviewed journals.

Purpose:
Artificial Production
Emphasis:
RM and E
Species Benefit:
Anadromous: 100.0%   Resident: 0.0%   Wildlife: 0.0%
Supports 2009 NPCC Program:
Yes
Subbasin Plan:
Fish Accords:
None
Biological Opinions:

Describe how you think your work relates to or implements regional documents including: the current Council’s 2014 Columbia River Basin Fish and Wildlife Program including subbasin plans, Council's 2017 Research Plan,  NOAA’s Recovery Plans, or regional plans. In your summary, it will be helpful for you to include page numbers from those documents; optional citation format).
Project Significance to Regional Programs: View instructions
The Hood River is specifically listed under RPA 64.1, projects directed at estimating the relative reproductive success of hatchery-origin salmon and steelhead. Our work is also very relevant to goal 64.2, “Determine if properly designed intervention programs using artificial production make a net positive contribution to recovery of listed populations”. That is because one needs a multi-generation pedigree to measure the effects of hatchery ancestry on the fitness of wild-born fish. In fact, the May 2010 document on Recommendations for Implementing RM&E for the 2008 NOAA fisheries FCRPS BiOp specifically recommends adding the Hood River steelhead program for coverage under RPA 64.2.
In this section describe the specific problem or need your proposal addresses. Describe the background, history, and location of the problem. If this proposal is addressing new problems or needs, identify the work components addressing these and distinguish these from ongoing/past work. For projects conducting research or monitoring, identify the management questions the work intends to address and include a short scientific literature review covering the most significant previous work related to these questions. The purpose of the literature review is to place the proposed research or restoration activity in the larger context by describing work that has been done, what is known, and what remains to be known. Cite references here but fully describe them on the key project personnel page.
Problem Statement: View instructions

 

Background: The current state of affairs

Hatchery fish have lower fitness than wild fish:

 Genetic change via  captive breeding has been documented in a wide variety of taxa, and it poses a serious problem for ex-situ conservation (captive breeding and release) programs (Frankham, 2008; Montgomery et al., 2010). This phenomenon is particularly well documented in salmonid fish.  Traditional hatchery stocks (those propagated for many generations using returning hatchery fish as parents) are genetically different from wild populations for a wide variety of adult and juvenile traits (reviewed in Fleming and Petersson, 2001, and in Bilby et al., 2005).  Fish from traditional hatcheries usually have much lower fitness than wild fish when they breed in the wild (often < 10% that of wild fish: reviewed in Berejikian and Ford, 2004, Araki et al., 2008, and Recovery Implementation Science Team, 2009). 

 What had not been appreciated until recently is how rapidly this fitness decline occurs.  For example, hatchery-produced Chinook salmon juveniles were less aggressive and more vulnerable to predators than wild juveniles after only a single generation in a state-of-the-art hatchery(Fritts et al., 2007; Pearsons et al., 2007).  Here the juveniles were raised in identical environments - they differed only in parentage.  Reisenbichler and McIntyre (1977) made all possible crosses among wild and hatchery steelhead trout (HxH, HxW, WxH and WxW), where the hatchery fish had been derived from the same wild population only two generations earlier.  Offspring were released into fenced off streams as eyed-embryos or fry, and recaptured 12 months later (parentage assigned using allozymes).  Juveniles produced from hatchery parents showed approximately 85-90% the survival of juveniles produced using wild parents, while hybrids were intermediate (Reisenbichler and McIntyre, 1977; Reisenbichler and Rubin, 1999).  When the same families were reared in a hatchery pond, the HxH offspring survived and grew best.

Consistent with the above results, three studies have now shown first-generation hatchery fish to have significantly lower lifetime reproductive success than wild fish (from two populations of steelhead and a population of coho).  E. Berntson and colleagues found the relative reproductive success (RRS) of first-generation steelhead in Little Sheep Creek, Oregon, to be consistently between 0.3 and 0.6 that of wild fish in six run years (E. Berntson, in preparation).  In a similar study, V. Theriault and colleagues studied three run years of coho in Calapooya creek, a tributary of the Umpqua River.  The RRS for first-generation hatchery fish averaged 0.56 to 0.77 that of wild fish (Theriault et al., unpub. data).  We found similar results for winter-run steelhead in the Hood River, as detailed below.

 We now have a three-generation pedigree of all the anadromous, winter-run steelhead (Oncorhynchus mykiss) that have spawned in the Hood River since 1991 (see study design and further details under “Major Accomplishments”). The older hatchery stocks of steelhead in the Hood River had 10-30% the fitness of wild fish when both spawn in the wild (Araki et al., 2007a).  First generation hatchery fish (i.e. fish created using wild parents as broodstock) have about 15% lower fitness than wild fish (Araki et al., 2007b).  A comparison between first and second-generation fish raised side-by-side in the same hatchery showed that each generation spent in the hatchery substantially reduced the performance in the wild of the hatchery fish (Araki et al., 2007b).  Because both types of fish experienced identical environments, and because the reciprocal crosses showed no evidence of maternal effects, the difference is almost certainly genetic.  We estimated an initial average decline in fitness of 37.5% per generation of hatchery rearing. 

 Data such as these are disturbing because hatchery and wild fish are allowed to mix freely in many rivers.  Furthermore, many hatcheries have adopted the mission of supplementation. The assumption had been that using local, wild-born fish as parents every year would circumvent the domestication effects that are so apparent in multi-generation stocks.  However, if that assumption is not true, then supplementation efforts may actually drag down the fitness of the wild stocks they are meant to support (Ford, 2002; Goodman, 2005; Lynch and O’Hely 2001).  Last year we analyzed the fitness of wild-born fish that differed in parentage, and found that wild-born offspring of naturally occurring HxH, WxH and WxW crosses also differed substantially in fitness (Araki et al., 2009).  Because the fish of different parentages presumably experienced identical environments (this time in the wild), the difference between them is probably genetically based.  Thus, these first-generation fish probably are dragging down the fitness of the wild population, although the extent of the effect still needs to be estimated accurately.

 

Objective I – finish and fully analyze the pedigrees

What still needs to be done?

In addition to improving our point estimates of RRS by adding more run years for the winter stock analyses, we can now test whether summer run Hnew differ from wild summers. This will be an independent test of the hypothesis that first-generation fish differ from wild.  However, one interesting difference between the summer-run and winter-run data sets is that the winter population had very little influence from any old hatchery stocks over the last two decades.  The summer run fish, on the other hand, had a  large influence from the highly domesticated Skamania stock, which returned to the Hood River by the thousands until the mid-1990’s. Therefore, the wild summer-run fish probably started with a big genetic handicap (i.e., high levels of hatchery introgression), relative to the wild winters.  So one interesting prediction is that there will be less of a fitness difference between wild and Hnew fish in the summer run than there was in the winter run. 

 In 1997 the last Skamania fish were passed over the dam and an Hnew stock of summer run fish was started.  In 1998-2000 only wild fish were deliberately let over the dam, and in 2001 the first of the Hnew summer stock began returning.  When we finish the 2009 run year of wild fish we will be able to estimate the fitnesses of each type of fish in run years 2001 to 2004 (>98% of offspring returned for RYs 01-03, and ~70% should have returned for RY 04).  One particularly interesting analysis we can do with the summer run is to estimate RRS of wild fish that differ in parentage, as we did for winter-run fish of wild and Hnew ancestry (Araki et al.2009).  However, here the hatchery influence is from the old Skamania stock, not an Hnew stock.  We estimate that in RY 2001 approximately 82% of the gene pool in the wild fish originated from Skamania fish.  That percentage drops to 57% in RY 2002, 21% in RY 2003 and 0% in RY 2004.  So it will be interesting to test whether the fitness of wild fish increases relative to that of the Hnew stock over the 2001-2004 period because the wild population will become “purer” (i.e. a smaller fraction of wild-born individuals descended from Skamania each year).  This analysis should give us an idea of how quickly a population rebounds once the influence of an old, out-of-basin stock is removed.   

Another question  we can now address  in both summer and winter programs is the effect of having individuals of mixed ancestry in the broodstock.  In other words, although a program may use only unclipped (wild-born) fish as broodstock, the managers cannot know the proportion of hatchery ancestry in those fish without an extensive pedigree.  One prediction is that wild-born broodstock with high hatchery ancestry will perform best in the hatchery (i.e. produce more surviving hatchery offspring), but that their offspring will perform worse than other hatchery fish in the wild.  If so, this effect would exacerbate the variance in performance among families both in the wild and in the hatchery.  In other words, broodstock fish that produce the most surviving offspring also produce the worst-performing offspring in the wild – a positive feedback that would enhance the negative effect of the hatchery on the wild population.  We have two pieces of data in support of this hypothesis.  First, the ratio of variance in family size (Vk) to mean family size (k) among wild fish in the Hood River was higher than typical values reported for other salmonid populations (Araki et al., 2007c).  We attribute this high variance to the presence of fish of varying hatchery background.  Second, we recently performed  the same analysis for broodstock used in the hatchery, and found a similarly large index of variability measured at their returning adult offspring.  In fact, the ratio of effective number of breeders (Nb) to actual number of breeders (Nc, for “census”) among hatchery broodstock was ~  0.50 across each of several brood years – not much higher than what we observe in wild fish.  Furthermore, egg take per female is not correlated with the number of offspring that return.  Figures 1 and 2 illustrate the high variability among broodstock pairs in their production of returning offspring.

Figure 1.  Distribution of number of returning offspring per hatchery broodstock pair (winter-run Hnew program) over nine brood years.  The distribution is as skewed as one observes in wild populations, with most pairs producing fewer than 5 returning offspring, and a few pairs producing dozens.  We suspect that selection is involved and will test whether the background of the broodstock parents explains some of this variance.

image001

Figure 2.  Principal coordinates analysis illustrating drift in the winter-run Hnew hatchery fish each year that results from the skewed (non-random) success among broodstock pairs in the hatchery.  Each point represents the allele frequencies in either the wild population (dark triangles) or in the returning hatchery fish (open triangles; several hundred fish each year) in each of 11 run years. The wild population has very constant allele frequencies from year to year, while there is a large amount of drift in the hatchery offspring each year.  Note that allele frequencies in the broodstock parents of those hatchery fish are indistinguishable from those in the wild (data not shown), so the drift results from the reproductive skew illustrated in figure 1 above.

image002

This was a well-run, conservation hatchery program, so the variance in family sizes shouldn’t be so large.  We therefore suspect that some sort of selection is involved.  As part of our continued analysis of the pedigree, we will test whether the ancestry of the broodstock explains some of that variance, and will also test for a negative correlation between performance in the hatchery (production of returning offspring) and performance in the wild (RRS of those returning offspring).  We will also explore phenotypic correlates of those two types of performance, including body size, run timing, cross date and MHC genotype.  An added twist to the analysis is that some fish will have resident trout for parents, and so we will test whether a resident parent is essentially equivilant to a wild anadromous parent, or has other effects.

  In summary, our first objective will be to complete the 19 year pedigrees for both winter  summer run fish.  We will analyze both pedigrees to tease apart the effects of hatchery ancestry on performance in the wild and in the hatchery. At the end of the next grant cycle we will have points estimates of RRS for Hnew vs. wild for multiple run years in each run, multiple estimates of the fitness of wild-born fish as a function of their ancestry (wild, Hnew, Skamania, and resident parentage), and an analysis of whether ancestry is responsible for the large skew in performance of broodstock.  At that point we will have completed the study that relied on complete samples of returning offspring each year.

 Because the Powerdale dam has been removed, all future pedigree work on the Hood River will have to use samples collected at weirs in particular tributaries in the Hood River (the ODFW will begin sampling weirs this year; R. French, pers. comm.).  Possible questions we might address with these samples include the effects of increased numbers of hatchery fish that will now be able to travel upstream (previously controlled at the dam),  including hatchery strays from other rivers (which could be substantial; Keefer et al., 2008).  We may also examine whether there is a genetic component to where fish go to spawn, and perform more in-depth analyses of interactions between anadromous and resident fish in particular tributaries.  For now, we will archive the samples until we have enough run years in hand to ask some interesting questions.

 

Objective II: Mechanisms – the next big question

 It is now clear that something about hatcheries causes a rapid and profound genetic change in steelhead, and probably in salmonids in general.  Unfortunately, we do not yet understand the mechanisms by which captive rearing causes such precipitous and genetically-based declines in fitness. 

 Why is it important to figure out the mechanism causing low fitness of hatchery fish? Given what we now know about the fitness of even first-generation hatchery fish, managers are now in a quandary given the competing goals of protecting native populations and insuring sufficient production for recreational and commercial fisheries.  One solution would be to physically segregate wild and hatchery fish.  This would be difficult and might require shutting down some hatcheries.  Alternately, one could move towards integrated broodstock programs and let the hatchery fish continue to mix with wild fish. The problem here is that even though first-generation fish (Hnew) are much more fit than old hatchery stocks, there is much more gene flow between Hnew fish and wild fish than between old stocks and wild fish.  So given they are still less fit than wild fish, the overall impact of Hnew fish on wild populations could still be important (as we showed in the Hood River).  Alternately, one could deliberately select for highly domesticated hatchery stocks (for example, stocks with very early run times) on the assumption that these changes would result in minimal gene flow into the wild populations with which they mix.  Difficulties associated with this strategy include unintended selection on correlated traits and some continued gene flow into wild populations, and fish that may be of less economic and societal value.  Each of these solutions obviously has its problems. 

 One way out of this situation would be to figure out how to make hatchery fish that don’t impose a genetic load on wild populations.  Thus, identifying what traits are under selection is a critical area of research.  Hypotheses for why hatchery fish are less fit than wild fish in nature include the following:

 (1) Relaxed natural selection:  Survival from egg to smolt is usually 85-95% in hatcheries versus 1-5% in the wild (Reisenbichler et al., 2004).  Thus, relaxed purifying selection during the egg-to-smolt stage could result in the accumulation of mutations that are neutral in the hatchery but harmful in the wild (Lynch and O’Hely, 2001).  Similarly, because adults are spawned artificially in the hatchery, there can be relaxed natural selection on adult reproductive traits.  For example, hatchery breeders do not choose or compete for mates, find suitable spawning sites, defend nests, and so on.  Nevertheless, relaxed natural selection seems an unlikely explanation for such dramatic declines after just a single generation of hatchery culture unless salmon carry a very high standing genetic load at many traits (e.g. as appears to be the case in oysters; Launey and Hedgecock, 2001).  Because typical rates of mutation to deleterious mutations are around one mutation per genome per generation, and the average effect of such a mutation in the heterozygous state is around 2% (Lynch et al., 1999), it should take at least a few generations for mutational effects to show up.  There is no evidence that salmon have unusually high mutation rates (e.g. rates at neutral microsatellite loci appear to be typical; Steinberg et al., 2002).  So unless the hatchery environment somehow induces a huge increase in the deleterious mutation rate, relaxed natural selection seems an unlikely explanation for such rapid declines in the first generation of hatchery culture, at least for typical traits under stabilizing selection.  On the other hand, for unusual traits that have a high standing genetic load, a single generation of relaxed selection might have an effect.  Such traits include those under cyclical selection, such as immune relevant genes (Hedrick, 2002; Tennessen and Blouin, 2008).  Under this scenario, the population can never optimally adapt to an environment that is a constantly moving target – in this case, pathogens.  Thus, part of what we propose below is to ask whether relaxed selection on mate choice causes hatchery fish to have unfavorable genotypes at the major histocompatibility loci (MHC), which are known to be under sexual selection and to influence disease resistance in salmon (more below).  We propose estimating the fitness costs of having suboptimal MHC genotypes, and estimating the gain in hatchery fish performance that could be achieved by choosing broodstock based on MHC genotype.

(2) Strong domestication selection: This is the traditional explanation for hatchery fitness declines.  Here positive selection for adaptation to the hatchery environment comes at the expense of adaptation to the wild environment (Ford, 2002).  A theoretical analysis by Araki et al. (2008) showed that strong selection on phenotypic variation generated among fish by the hatchery is indeed sufficient to get fitness declines of the magnitude observed in the Hood River.  The fact that after just one or two generations hatchery fish perform better than wild fish in hatchery environments (Reisenbichler and Rubin, 1999) also points to selection, rather than to some generalized genomic deterioration as the source of genetic change in hatchery fish.  If we find a negative correlation between performance in the hatchery and performance in the wild (as outlined above), then that would be further support for the hypothesis of domestication selection.  What traits might be under strong directional selection remains a mystery, but a plausible case has been made for growth rate (Reisenbichler et al., 2004).  We are currently testing that hypothesis by comparing the growth of fish of hatchery and wild backgrounds raised in a common garden experiment.  Below we also propose a novel approach for identifying traits that were under selection in the hatchery: searching for genes that are differentially expressed between fish of hatchery and wild background.

 (3) More exotic mechanisms: Two intriguing possibilities are: (1) the hatchery induces a high mutation rate in the offspring (which then makes the mutation accumulation hypothesis more plausible), and (2) heritable epigenetic changes that are induced by the hatchery environment.  Epigenetic changes are mitotically or meiotically heritable changes in gene expression that result from modifications to the DNA without change in the DNA sequence (Richards, 2006; Bird, 2007; Jirtle and Skinner, 2007).  Most importantly, epigenetic changes can be transmitted to the next generation (Reik 2007; Richards, 2006; Jirtle and Skinner 2007).  It is therefore conceivable that rearing in a hatchery environment during the early part of the salmon lifecycle could cause epigenetic changes that eventually affect the fitness of those individuals and their offspring.  We recently showed that hatchery and wild adult steelhead from the Hood River do not differ in overall levels of genomic methylation (Blouin et al., 2010).   However, altered methylation of of a few DNA sites or other epigenetic processes (e.g. histone modifications) could still be important, so the epigenetics hypothesis deserves further study.  Rather than propose further studies at this time on genome-wide epigenetic differences between types of fish, we will be guided by  the results of the gene expression studies proposed below.  Epigenetic modifications exist to regulate gene expression, so it is conceivable that differential expression we observe for some genes might actually result from environmentally induced epigenetic effects rather than traditional selection.  We may return to this hypothesis in the future, depending on the outcome of the gene expression studies (see below).

 One final comment to put the work on mechanisms into context:  Whatever is going on in hatcheries almost certainly involves multiple traits, with some under directional selection and some under relaxed selection (e.g. Araki et al., 2008).  Thus, a multipronged approach to modifying hatchery practice will be necessary.  Here we are asking (1) can we identify particular traits that were the targets of directional selection, and (2) could reversing the relaxed selection on immunity genes contribute significantly as part of that multipronged approach.

 

Objective II, Deliverable 1. Gene expression study

Why look at gene expression?

There are two approaches to figuring out which traits are under selection.  First, one can make educated guesses and then test those candidate traits (the “top down” approach).  For example, there is a plausible argument that hatcheries select for excessively fast growth rate, and we are currently testing that hypothesis.  However, many other traits could be involved, such as adult breeding behaviors, sperm competitive abilities, adaptation to water temperatures or crowding levels that differ from those in the wild, and so on.  There may be traits we haven’t even considered.

 The second, “bottom-up”, approach to figuring out which traits are under selection is to figure out what changed in the genome of hatchery fish, and to then ask what traits those genes control.  In the first attempt of this kind, Roberge et al. (2006) used microarrays to perform gene expression profiling in two stocks of farmed Atlantic salmon (5-7 generations in captivity) and in their wild progenitor populations.  They found many differentially-expressed genes, including seven that had changed in parallel in the two farmed stocks.  Several of these genes were in growth-regulating pathways, but several were novel and may point to unknown traits under selection.  Indeed, gene expression studies frequently reveal new and unexpected genetic pathways involved in phenotypic variation (Whitehead and Crawford, 2006). 

 Finding the genes that responded to the hatchery environment  should then point us to potential traits under selection.  That knowledge will help us decide how to modify hatchery practices to reduce those selection pressures (it is also possible that certain genes are differentially expressed, not because of traditional directional selection, but because of heritable, environmentally-induced epigenetic effects.  We may return to this hypothesis in the future).

 Here we will use the gene expression profiling approach to identify potential traits that differ genetically between hatchery and wild fish.  However, we will use a new technology for gene expression analysis that is an improvement over the traditional microarray approaches.  We will use the Illumina next generation sequencing technology available at OSU (see below) to quantify expression in the entire transcriptomes of individual fry that differ only in parentage (first-generation hatchery parents versus wild parents).  We will compare fish reared in the same environment, but whose parents were either wild or first-generation hatchery fish, and do this for stocks from two independent river systems.  Response by the same genes in two independent hatchery stocks would be strong evidence for parallel evolution under similar selective pressures (Derome et al., 2006; Roberge et al., 2006). 

Gene expression analysis involves comparing the extent to which individual genes are “turned on” by counting the number of times we observe messenger RNAs (mRNA) transcribed from each gene.  Messenger RNA transcription level (“expression”) is highly correlated with protein production.  Thus, changes in transcription rate at individual genes are a major way in which differences in phenotype (traits) evolve among individuals.  Thus, it is most likely that the genetic change in hatchery fish involves increased or decreased transcription at some key genes.  So this is a reasonable  first place to look. 

 Expression profiling involves counting the number of times we see mRNA from each gene in a tissue sample.  Here we will use a new method for counting these mRNAs, which involves sequencing the transcriptome (i.e. all the mRNA in the sample) of each fish (the mRNA-seq procedure; Mortazavi et al., 2008) (http://en.wikipedia.org/wiki/RNA-Seq).  We extract total mRNA from a fish, convert it into cDNA, and then sequence that cDNA library using the new, high-throughput sequencing technologies that are now available.  The result is a digital count of the number of times each gene was represented in the complete mRNA sample.  The identity of each read (what gene it came from) is deduced by comparing your reads to a reference genome for the species or, for species with out a complete genome,to a reference transcriptome (all the known coding regions). These counts are analyzed by ANOVA with corrections for false discovery rate to decide if your types of fish (e.g. H vs. W) differ in expression for a particular locus.  OSU’s Center for Genome Research and Biocomputing has an Illumina 1G paired-end sequencing system that we use for this procedure (http://www.illumina.com/technology/mrna_seq.ilmn).  The sample preparations and bioinformatics analyses will be conducted in collaboration with my colleague at OSU, Dr. Todd Mockler, who specializes in Illumina-based expression profiling and downstream gene ontology analysis (e.g.  Filichkin et al., 2009; Fox et al., 2009) (see letter of support by Mockler in "Project Documents"). 

 The samples to use are already in hand:

We froze in liquid nitrogen the offspring from 15 HxH and 17 WxW fish families created from Hood River fish in 2009 (~12 offspring per family frozen just after yolk absorption). In 2010 we created and froze offspring from seventeen 2x2 matrices, each created by crossing an H and W male with an H and W female (siblings of fish used in the growth rate experiments described under “Accomplishments”).  The reciprocal “hybrids” in this design will allow us to separate additive genetic from non-additive and maternal effects.  These samples are all stored at -80 Celsius and await processing.  We also have similar samples and sample sizes created using wild and first-generation hatchery steelhead from the Siletz River, which we will use to see if the same selection pressures operated in independent hatcheries.

 Why look at gene expression in fry, as opposed to other life stages?

We froze the fry just after yolk absorption, when they would begin feeding.  This is a major transition period in development and a time of high mortality in the wild. Also, evidence points to differential survival in the wild among juveniles as the key life stage in which genetic differences result in fitness differences.  For example, studies have shown reduced survival of hatchery fish during the egg to parr or egg to 1-year old stages (Reisenbichler and McIntyre, 1977; Reisenbichler and Rubin, 1999).  Several studies have demonstrated behavioral differences between juveniles of hatchery and wild origin (e.g. aggression, vulnerability to predators; Fritts et al., 2007; Pearsons et al., 2007).  Therefore, we hypothesize that gene expression differences are most likely to be expressed in the early juvenile stages.  This is also the stage analyzed by Roberge et al. (2006) who found consistent gene expression differences in two stocks of farmed salmon and their wild progenitors.  In the future we may follow this study by looking at expresson differences in other life stages.

 Preliminary data and progress so far:

The genome of O. mykiss has not yet been sequenced.  Therefore, the first step in this project was to sequence Hood River fish transcriptomes very deeply and with extra-long reads, and then align those reads with the genomes of the five fish species for which we do have completed sequences.  This allows us to build a reference transcriptome for O. mykiss that we can use for the actual expression analysis of individual fish (each fish sequenced much less deeply than the ones used to build the reference).  We successfully sequenced four Hood River fish with 80 bp paired-end runs, each to a depth that would average 10X coverage of all genes assuming a ~30,000 gene genome and typical distribution expression differences among genes. We are now assembling the reference transcriptome.  However, we have recently learned that a complete genome for O. mykiss is in progress  and may be available in around a year’s time (M. Miller, Univ. Oregon, pers. comm.).  When the complete genome  is available, we can re-analyze the data we are now generating to see if we can identify more of our reads (a certain percentage are always binned as ‘unknown’) and identify more of the rarer transcripts.

 We currently have a small grant ($30,000) from the Oregon Department of Fisheries and Wildlife for supplies to complete preliminary work on this question. That money allows us to build the reference transcriptome and should provide for running one or two fish from each of 30 families (15 HxH and 15 WxW).  We just finished preparing our first 8 expression samples and will be doing preliminary runs this month to optimize the number of samples to multiplex per lane.

 Here we are requesting funding from BPA to perform this project in more depth, including both years of crosses and, eventually, more fish per family.  We also would like to test whether loci identified as differentially expressed in the Hood River also differ between fish of H and W background from the Siletz River (for this latter step we would use quantitative PCR to test the most interesting individual genes).  We expect this work to take about 2 years, and propose picking it up again in the first year of this project.  By that time we’ll have finished and analyzed the initial samples funded by ODFW, and we hope, have an even better reference transcriptome in the form of a completed O. mykiss genome.  Also, the next generation sequencing technology is advancing so rapidly that in two years we’ll probably be able to examine more individuals at lower cost.

 Current estimates of cost:

Based on our preliminary runs, we calculate the costs of reagents for sample preparation and the fees for using the Illumina to currently be approximately $500 per fish.  This includes: Illumina mRNA-seq kit (RS-100-0801) = 8 mRNA-Seq samples for $1,940.00

= $242.5 per fish for the sample prep. Paired-end 40 cycle sequencing = $1,044 per lane (includes sequencing reagents, backup tapes, central services lab fees) assuming we 4-plex in each lane = $261/fish for the sequencing.  If we do one fish per family from each of the seventeen 2x2 matrices we created in 2010, that would be 68 fish for $34,000 at today’s prices (which we, again, expect to come down substantially).  With the ODFW-funded samples from 2009, this would give us unrelated fry from a minimum of 32 HxH families and 32 from WxW families, and 17 each from HxW and WxH families.  Any strongly up or down regulated genes between fish types should be apparent with those sample sizes.

 

Objective II, Deliverable 2. Traits involved in relaxed natural selection – MHC?

As mentioned above, the effect of relaxed selection should not be evident after just one or two generations of hatchery culture unless the there is a high standing genetic load (many suboptimal alleles segregating) for the traits involved. Such a high genetic load seems implausible for most traits (which should be under stabilizing selection) but could be true for loci under cyclical or fluctuating selection (Hedrick, 2002; Tennessen and Blouin, 2008).  In the case of immunity genes, temporally fluctuating selection can also act in concert with traditional frequency-dependent selection, in which rarity of alleles per se is favored (Hedrick, 2002).  Now consider the choice of a small number of wild broodstock for a conservation hatchery.  If choice is random, then first, one is likely to get the most common alleles (often the least fit) in the broodstock, and second, there is no opportunity for sexual selection to sort among those alleles. All individuals get to breed, no matter their allelic status. Also, the lack of mate choice means there is no opportunity for creating the most favorable combinations of alleles in heterozygous offspring.

 Loci of the major histocompatibility complex (MHC) are an essential part of vertebrate immune systems.  They are among the most polymorphic coding genes known, and have been shown to be under strong selection in many species, including salmon and trout (reviewed in Bernatchez and Landry, 2003). It is well documented that females of many species prefer males with different MHC genotypes (e.g. Penn, 2002; Bernatchez and Landry, 2003; Mays and Hill, 2004; Ziegler et al., 2005), as heterozygous offspring tend to be fitter than homozygotes, and heterozygotes for the most dissimilar genes have the greatest fitness.  For example, non-random mating with regard to MHC genotypes has been demonstrated in Chinook salmon (Neff et al., 2008), brown trout (Forsberg et al., 2007) and Atlantic salmon (Landry et al., 2001).  It is also well established in many species that MHC genotype influences disease resistance, with many examples found in salmonid fish (reviews in Bernatchez and Landry, 2003, and Spurgin and Richardson, 2010).  For example, MHC genotype controls resistance to infectious hematopoietic necrosis virus (IHNV) in chinook salmon (Arkush et al., 2002), to furunculosis (Aeromonas salmonicida) and infectious salmon anemia virus (ISAV) in Atlantic salmon (Lohm et al., 2002; Grimholtz et al., 2003), to Pseudomonas fluorescens in whitefish (Wedekind et al., 2004), and to bacterial coldwater disease (Flavobacterium psychrophilum) in O. mykiss (Johnson et al., 2008), which is our study species.  In O. mykiss it was also observed the IHNV upregulates expression of certain MHC loci (Hansen and Patra, 2002).  Note that IHNV is a pathogen of concern in the Hood River, and one for which all broodstock are tested (C. Banner, ODFW). Lastly, MHC loci in salmon and steelhead show patterns of allele frequency divergence among populations that differ from those at neutral loci (e.g. Evans et al., 2010; Aguilar and Garza, 2006), which further suggests that these loci are under selection.

 Relaxed selection by hatcheries on which fish get to reproduce could result in major fitness effects at MHC loci.  In one compelling example from Atlantic salmon  (Consuegra and Garcia de Leaniz, 2008), MHC alleles appeared to be paired randomly in local-origin, first-generation hatchery fish, but alleles in heterozygotes were more dissimilar than expected by chance in their wild cousins (which implies disassortative mating or non-random survival with respect to MHC).  Furthermore, MHC allele dissimilarity was negatively correlated with parasite levels within each type of fish, and hatchery fish were more heavily parasitized than wild.  In another example, which used a lab-reared factorial cross among 11 male and 11 female Chinook salmon of known MHC genotype, Pitcher and Neff (2007) observed substantial non-additive effects on offspring survival and growth to 80 days post fertilization.  They estimated that pairing adults assortatively by MHC would have increased offspring performance by 6% - and this was just to the fry stage.  The payoff over the entire life cycle could be much larger.

 Thus, a great deal of evidence suggests that either letting broodstock choose their own mates (not very practical), or having managers choose broodstock based on MHC as a proxy might give an immediate payoff by creating fitter first-generation hatchery fish.  Whether that benefit would extend into the second, natural born generation depends on whether the allelic effects are mainly additive (some alleles are better than others, on average) or non-additive (it is the particular pair of alleles in a heterozygote that matters) (“good genes “ vs. “compatible genes” in the terminology of Neff and Pitcher, 2005).  Evidence for each type of effect runs about 50:50 in the literature.  For example, only additive effects of particular MHC alleles are observed in the Atlantic-salmon-furunculosis system (Lohm et al. 2002), while Consuegra and Garcia de Leaniz (2008) observed only non-additive effects.  If the effects are mainly additive, then they would indeed carry over to the next generation.  If effects are mostly non-additive, then pairing parents disassortatively by MHC would get you a boost in fitness of the first-generation hatchery fish, but any effect would vanish in the next generation as those allelic combinations are broken up.

 We propose testing whether MHC influences the fitness of hatchery and wild fish in the Hood River.  Here we want to know if there is any evidence to support the idea that choosing broodstock based on MHC and/or deliberately mating MHC-dissimilar males and females would (1) improve hatchery production (i.e. produce hatchery fish that grow better and are more likely to return as adults), and (2) produce hatchery fish that are more like wild fish (i.e. have better reproductive success in the wild, given they survive to return as adults).  A direct, experimental test of these hypotheses would take over a decade to complete. However, we have already estimated the lifetime fitness of thousands of fish of varying hatchery and wild background, including the performance of hundreds of fish as broodstock in the hatchery.  Therefore, we simply have to sequence MHC loci in a selected set of those fish to answer the following questions. 

 

 For fish arriving at the dam:

1) Is the RRS of steelhead in the Hood River correlated with their MHC genotypes, and if so, do the effects appear to be additive or non-additive?

 2)  Do hatchery fish wind up, on average, with less favorable MHC genotypes than wild fish?

 

 For matings conducted in the hatchery:

1) Did broodstock pairs having either better individual alleles (additive effect), or better matched alleles (non-additive effect) produce more returning hatchery offspring (here “better” is as inferred from 1 & 2 above)?

 2) Within full-sib families, were the survivors a non-random subset of the possible sibling genotypes?

 

 Direct tests of disease resistance:

We also propose doing two direct tests of the effects of MHC on disease resistance in Hood River fish.  These include:

 1) We will also sequence MHC in the fish used for the disease challenge experiments described in the “Accomplishments” section.  Here we will test if MHC influences survival differences among sibships, or among siblings of different genotype within sibships.

 2) We propose harvesting 200 returning winter-run hatchery fish and quantifying their parasite loads (we can’t use wild fish because the population is federally protected).  Then we will ask if MHC genotypes correlate with their naturally-acquired parasite load (as in Consuegra and Garcia de Leaniz, 2008).  These fish will be collected by ODFW staff at a weir that they will run at the base of the East fork of the Hood River for the purpose of keeping hatchery fish out of that fork.  The parasitology work will be done by Kym Jacobson’s lab (NOAA Fisheries, Hatfield Marine lab; e.g. Baldwin et al., 2008; Jacobson et al., 2008)(see letter of support in "Project Documents"). 

 

 Methods

Classical MHC loci are involved in distinguishing foreign from ‘self’ molecules, and in identifying pathogens for destruction.  Class I MHC loci recognize intracellular pathogens, while class II MHC loci recognize extracelluar pathogens.  Given that both class I and class II loci have been demonstrated to be under fluctuating and directional selection, yet target different types of pathogens, we will sequence individuals at both types of loci.  Furthermore, because there are two exons (α and β) associated with class II MHC loci that exhibit different levels of polymorphism, we will sequence individuals at both of these loci.  There are multiple loci associated with class I MHC in O. mykiss. We will focus on MHC Iα, which is most commonly associated with disease resistance (Dijkstra et al. 2007).  Importantly, the class II MHC loci have no paralogues (homologous gene copies to which the primers could bind) (Gomez et al. 2010), which means that we can reliably amplify our region of interest. 

 We propose sequencing 450 fish from the disease-challenge trails, the 200 returning adult fish collected for parasite surveys, and 2000 fish from the main pedigree (including broodstock) to answer the questions posed above.  Each will be sequenced at the MHCIα, MHCIIα and MHCIIβ loci.

 

Preliminary data

To validate our proposed methodology, we sequenced 30 winter-run, Hood River steelhead at the MHCIα, MHCIIα and MHCIIβ loci.  Primers for these three loci were identified from previous studies (Miller et al.1997, Gomez et al. 2010) and verified by matching primer sequences to the curated collection of O. mykiss MHC sequences found at the immuno-polymorphism database (http://www.ebi.ac.uk/ipd/mhc/fish/index.html).  All loci give clear, consistently readable electropherograms (e.g. Fig. 3), demonstrating that these loci and primer pairs will be suitable for the proposed study.  Note also that the PCR works equally well on DNA from fin snips and on our oldest scales from the early 1990’s.  Therefore we will be able to use the entire pedigree for this study.  We identified 19 polymorphic sites with 10 nonsynonymous substitutions at the MHCIIβ locus.  Interestingly, at the MHCIα locus, each sequenced individual had at least one unique allele, which is not surprising given the high polymorphism observed at this locus (Dijkstra et al. 2007).  MHCIIα was the least polymorphic locus, with 8 polymorphic sites comprising 3 unique alleles.  The methods work well in our lab and we already have  the necessary samples in hand, which makes us well prepared to complete this project.  Furthermore, as noted above, our extensive pedigree data set uniquely positions us to address these important questions. 

 Figure 3. Electropherograms of MHCIIβ sequence data from five Hood River steelhead. 

image003

 

Timeline

Here we request four years of funding.  By that time we should have answered all the key questions we originally set out to answer in terms of RRS of summer and winter fish in the Hood River. For the future I think it is most interesting to focus on the mechanisms of fitness decline.  For now we have four main datasets to complete over four years: finish and analyze the summer-run fish, finish and analyze the winter-run fish, do the gene expression study, and do the MHC study.  We could start with any of them in 2012, but in order to keep the yearly budgets at about the same amount as in previous years, I propose spreading the four projects out as in the figure below.  In the “Deliverables and budget” section I assigned about a quarter of the total four-year budget to each of these four main datasets, which is probably pretty close to what each will cost.  In FY 2011, which is the last year of our current funding, we will focus on the summer-runs and the growth rate and disease challenge experiments (see “Accomplishments” for more info on those later two in progress). We will continue in the new funding cycle with the goal of completely finishing the summer run dataset in fy 2013.  Then, we will return to the winter run and finish up that analysis by 2015 (we have already published extensively on the winter run, and so want to ask the same questions on the summers before returning to the winter run).  We’ll  perform the gene expression study in 2012-2014, and the MHC study in 2014-2015.  Doing the MHC later makes sense because by then we’ll have estimated lifetime fitness for more of the summer fish. 

We have not requested a fifth year of funding at this point because it is hard to predict what the key questions will be that far out.  It largely depends on what we find out in the next few years.  For example, we hope to identify some key genes that differ in expression between fish of H and W background.  If so, we would then want to design tests of hypotheses about the traits those genes control.  We may also return to the epigenetics hypothesis.  Finally, by then we will have accumulated five years of samples from the weirs that will be operated by ODFW in the Hood River, and could ask several interesting questions using those samples (e.g. there may be a large influx of strays now that the dam is gone; Keefer et al., 2008.  What effect might they have?).  

image004


What are the ultimate ecological objectives of your project?

Examples include:

Monitoring the status and trend of the spawner abundance of a salmonid population; Increasing harvest; Restoring or protecting a certain population; or Maintaining species diversity. A Project Objective should provide a biological and/or physical habitat benchmark by which results can be evaluated. Objectives should be stated in terms of desired outcomes, rather than as statements of methods and work elements (tasks). In addition, define the success criteria by which you will determine if you have met your objectives. Later, you will be asked to link these Objectives to Deliverables and Work Elements.
Objectives: View instructions
Describe fitness effects of hatchery summer run steelhead (OBJ-1)
We propose completing the full 19-year pedigree for summer run fish in the Hood River, and then analyzing the fitness of fish in the wild and in the hatchery (when used as broodstock) as a function of their ancestry. Here ancestry includes new and old stocks of hatchery fish, wild anadromous and resident fish.

Describe fitness effects of hatchery winter-run steelhead (OBJ-2)
We propose completing the full 19-year pedigree for winter run fish in the Hood River, and then analyzing the fitness of fish in the wild and in the hatchery (when used as broodstock) as a function of their ancestry. Here ancestry includes mainly a new stock of hatchery fish, wild anadromous and resident fish.

Identify genes differentially expressed between fry of H and W ancestry (OBJ-3)
Here we will scan the genomes of juvenile fish of wild and hatchery ancestry (raised in the same environment) for genes that differ in expression levels. Then we will analyze the physiological pathways in which those genes are involved, with the ultimate goal of identifying traits that were under selection in the hatchery.

Test effects of relaxed selection on immunity genes (MHC) (OBJ-4)
Here we want to know if relaxed selection on MHC genes could explain a significant amount of the fitness drop in hatchery fish, and if incorporating MHC genotype into criteria for choosing broodstock could result in a significant improvement in the fitness of hatchery fish.


The table content is updated frequently and thus contains more recent information than what was in the original proposal reviewed by ISRP and Council.

Summary of Budgets

To view all expenditures for all fiscal years, click "Project Exp. by FY"

To see more detailed project budget information, please visit the "Project Budget" page

Expense SOY Budget Working Budget Expenditures *
FY2019 $301,526 $301,526 $337,434

BiOp FCRPS 2008 (non-Accord) $301,526 $337,434
FY2020 $150,763 $175,000 $164,892

BiOp FCRPS 2008 (non-Accord) $175,000 $164,892
FY2021 $100,000 $125,000 $123,023

BiOp FCRPS 2008 (non-Accord) $125,000 $123,023
FY2022 $0 $12,342

BiOp FCRPS 2008 (non-Accord) $0 $12,342
FY2023 $0 $0

FY2024 $0 $0

FY2025 $0 $0

* Expenditures data includes accruals and are based on data through 31-Mar-2025

Actual Project Cost Share

The table content is updated frequently and thus contains more recent information than what was in the original proposal reviewed by ISRP and Council.

Current Fiscal Year — 2025
Cost Share Partner Total Proposed Contribution Total Confirmed Contribution
There are no project cost share contributions to show.
Previous Fiscal Years
Fiscal Year Total Contributions % of Budget
2021
2020
2019
2018
2017 $9,020 3%
2016 $18,895 5%
2015
2014
2013
2012
2011
2010
2009
2008
2007

Discuss your project's recent Financial performance shown above. Please explain any significant differences between your Working Budget, Contracted Amount and Expenditures. If Confirmed Cost Share Contributions are significantly different than Proposed cost share contributions, please explain.
Explanation of Recent Financial Performance: View instructions
We have always spent at least 98% of the funds allocated to us each year, and we have never gone over budget.
Discuss your project's historical financial performance, going back to its inception. Include a brief recap of your project's expenditures by fiscal year. If appropriate discuss this in the context of your project's various phases.
Explanation of Financial History: View instructions
In the "summary from recent years" table above, the contracted amounts are not listed except for in the current FY, 2010. The numbers in the expenditures column do not match our contracted amounts or expenditures per fiscal year, nor do they even sum to the same amounts. So I don’t understand the origin of those figures. The true amounts each year can be seen in the ‘Fiscal Year Budgets’ pdf in the “Budgets” link to the upper left of the above table. The numbers in that pdf are correct. Please note two potentially confusing items in the detailed expenditures pdf: (1) The FY06 budget is actually the sum of FY05 and FY06. (2) In FY09 we were given a budget of only $235,177. It now appears that at some point an additional $41,823 was approved to bring that year’s budget up to the BPA approved expense budget of $277,000. However, I was never informed of the availability of that additional funding. That extra $41,823 now appears as an unallocated expense budget item under FY09, which makes it appear as if I under spent my budget by 15% that year. Spending in context of the project’s various phases: The distribution of effort between lab work and data analysis has remained pretty constant throughout the phases of this project, and should continue so into the future.

Annual Progress Reports
Expected (since FY2004):17
Completed:15
On time:15
Status Reports
Completed:70
On time:32
Avg Days Late:1

                Count of Contract Deliverables
Earliest Contract Subsequent Contracts Title Contractor Earliest Start Latest End Latest Status Accepted Reports Complete Green Yellow Red Total % Green and Complete Canceled
19502 PI 200305400 REPRO OF STEELHEAD IN HOOD RIV Oregon State University 10/01/2004 09/30/2006 Closed 4 5 0 0 0 5 100.00% 0
29562 35294, 39163, 43555, 49886, 54599, 58865, 63427, 67127, 70177, 74233, 77259, 76914 REL 2, 76914 REL 4, 76914 REL 6 2003-054-00 EXP REPRO OF STEELHEAD IN HOOD RIV Oregon State University 10/01/2006 10/31/2021 Closed 66 100 0 0 7 107 93.46% 2
Project Totals 70 105 0 0 7 112 93.75% 2

Selected Contracted Deliverables in CBFish (2004 to present)

The contracted deliverables listed below have been selected by the proponent as demonstrative of this project's major accomplishments.

Contract WE Ref Contracted Deliverable Title Due Completed
19502 F: 161 Meeting attendance documented in final report 9/28/2006 9/28/2006
19502 B: 157 Successfully genotype approximately 3600 fish 9/28/2006 9/28/2006
19502 C: 162 Complete interpetation of data 9/28/2006 9/28/2006
19502 D: 132 Submit FY06 Annual Report 9/28/2006 9/28/2006
29562 D: 161 Meeting attendance documented in final report 9/30/2007 9/30/2007
29562 B: 162 Complete interpretation of data 9/30/2007 9/30/2007
29562 F: 132 Submit FY07 Annual Report 9/30/2007 9/30/2007
35294 D: 161 Meeting attendance documented in final report 9/30/2008 9/30/2008
35294 B: 157 Successfully genotype approximately 3000 fish 9/30/2008 9/30/2008
35294 C: 162 Complete interpretation of data 9/30/2008 9/30/2008
35294 F: 132 Attach FY08 Annual Report in Pisces 9/30/2008 9/30/2008
39163 D: 161 Meeting attendance documented in final report 9/30/2009 9/30/2009
39163 B: 157 Successfully genotype approximately 3000 fish 9/30/2009 9/30/2009
39163 C: 162 Complete interpretation of data 9/30/2009 9/30/2009
39163 F: 132 Attach FY09 Annual Report in Pisces 9/30/2009 9/30/2009

View full Project Summary report (lists all Contracted Deliverables and Quantitative Metrics)

Discuss your project's contracted deliverable history (from Pisces). If it has a high number of Red deliverables, please explain. Most projects will not have 100% completion of deliverables since most have at least one active ("Issued") or Pending contract. Also discuss your project's history in terms of providing timely Annual Progress Reports (aka Scientific/Technical reports) and Pisces Status Reports. If you think your contracted deliverable performance has been stellar, you can say that too.
Explanation of Performance: View instructions
Every year of this project we have produced all the promised deliverables, except in one year we did not genotype as many fish as promised owing to a personnel change that caused us to have to redo a large number of samples. Else we have consistently genotyped all the fish proposed, completed all the analyses promised, and communicated our results yearly by attending meetings and publishing promptly. Although I have sometimes been a few days late with status reports and annual reports, I have always produced the required documents.

  • Please do the following to help the ISRP and Council assess project performance:
  • List important activities and then report results.
  • List each objective and summarize accomplishments and results for each one, including the projects previous objectives. If the objectives were not met, were changed, or dropped, please explain why. For research projects, list hypotheses that have been and will be tested.
  • Whenever possible, describe results in terms of the quantifiable biological and physical habitat objectives of the Fish and Wildlife Program, i.e., benefit to fish and wildlife or to the ecosystems that sustain them. Include summary tables and graphs of key metrics showing trends. Summarize and cite (with links when available) your annual reports, peer reviewed papers, and other technical documents. If another project tracks physical habitat or biological information related to your project’s actions please summarize and expand on, as necessary, the results and evaluation conducted under that project that apply to your project, and cite that project briefly here and fully in the Relationships section below. Research or M&E projects that have existed for a significant period should, besides showing accumulated data, also present statistical analyses and conclusions based on those data. Also, summarize the project’s influence on resource management and other economic or social benefits. Expand as needed in the Adaptive Management section below. The ISRP will use this information in its Retrospective Review of prior year results. If your proposal is for continuation of work, your proposal should focus on updating this section. If yours is an umbrella project, click here for additional instructions. Clearly report the impacts of your project, what you have learned, not just what you did.
All Proposals: View instructions
  • For umbrella projects, the following information should also be included in this section:
  • a. Provide a list of project actions to date. Include background information on the recipients of funding, including organization name and mission, project cost, project title, location and short project summary, and implementation timeline.
  • b. Describe how the restoration actions were selected for implementation, the process and criteria used, and their relative rank. Were these the highest priority actions? If not, please explain why?
  • c. Describe the process to document progress toward meeting the program’s objectives in the implementation of the suite of projects to date. Describe this in terms of landscape-level improvements in limiting factors and response of the focal species.
  • d. Where are project results reported (e.g. Pisces, report repository, database)? Is progress toward program objectives tracked in a database, report, indicator, or other format? Can project data be incorporated into regional databases that may be of interest to other projects?
  • e. Who is responsible for the final reporting and data management?
  • f. Describe problems encountered, lessons learned, and any data collected, that will inform adaptive management or influence program priorities.
Umbrella Proposals: View instructions

 I. Main question: fitness of hatchery versus wild fish

 We compared the relative success of two "old" hatchery stocks vs. wild fish (the winter run “Big Creek” stock and the summer run “Skamania” stock), and showed they have much lower total fitness than wild fish when both breed in the wild (RRS ~ 0.10 for Big Creek, ~0.30 for Skamania; Araki et al., 2007a).  Note that the “wild” summer population had been swamped for years by Skamania fish (e.g. 3-8 times as many hatchery fish as wild fish were passed over the dam each years in 1992-1997.  So the fitness difference between Skamania and a truly wild population would probably have been even more drastic.  In that same paper we also concluded that the winter-run Hnew were not significantly different from wild fish, based on 3 run years of data.  But a subsequent analysis based on six run years of data showed the difference was significant, with the Hnew winter run fish averaging about 85% the fitness of wild fish (Araki et al., 2007b).  We now have several more run years in which we can add further point estimates, but the analysis becomes much more complex because fish of a variety of genetic backgrounds are returning (and also being used as broodstock).

 Fig. 1  Example: design to estimate relative reproductive success of Hnew vs. wild fish in winter run, 1995-96 run year. Circles = run years.  First generation of Hnew F1 hatchery fish were created and released in 1992.  They returned to spawn in the wild mostly in 1995 and 1996.  The offspring of the F1’s that spawned in 1995 are illustrated as an example.  % = fraction of that cohort that returned in each subsequent year.  From their F2 offspring that returned from 1998 to 2001, we estimated the fitness of each type of fish (Hnew vs. wild) that spawned in 1995. 

 

 

 

fig 1 again b 

In Araki et al. (2007b) we compared the first generation (Hnew) fish with second-generation hatchery fish (created using one wild parent and one Hnew parent) raised in the same hatchery (Fig. 2).  Second generation fish had ~55% the reproductive fitness of the first generation fish (Araki et al., 2007b).  There was no evidence for maternal effects and both types of fish experienced identical environments, so the difference between them must be genetically based.  This result suggests that the decline in fitness that results from recycling “hatchery genes” back through the hatchery can occur very quickly. 

Fig. 2  Design for estimating fitness of 1st vs 2nd generation hatchery fish.  (A) Experimental design for estimating the fitness (relative reproductive success, RRS) of first- verus second-generation hatchery fish (illustrates a single run year, 98-99, but we did this for 3 run years).  The dotted lines represent the returning, wild-born offspring used to estimate the fitness of the two types of hatchery fish.  (B) Reproductive success of each type of fish relative to wild, plotted against generations in captivity.  0 generations = wild fish, which have RRS = 1, by definition. 1 = first-generation hatchery fish (two wild parents), and 2 = second generation (one parent of each type).  We have analyzed six years of data in total for 1st generation (Hnew) fish, and 3 years for 2nd generation H fish  

fig 2A again 

Fig. 2B

image003

Thus, we demonstrated a genetically-based effect of hatchery culture that reduces fitness in the wild and that accumulates with each generation of hatchery culture.  Nevertheless, even if captive-bred individuals are genetically different and produce fewer offspring than wild individuals, adding them to a wild population can still give a demographic boost without substantial harm to a wild population that is below carrying capacity if (1) the genetic effects do not persist into the next generation (i.e., natural selection purges the offspring generation of their deleterious alleles before they reproduce), and (2) enough captive-bred individuals are added each generation to make up for their lower productivity.  If the first condition is not true, however, genetic effects will accumulate over time, potentially leading to a downward spiral in the absolute fitness of the supplemented wild population (Lynch and O’Hely, 2001).  Thus, the key question is whether the wild-born descendents of captive-bred fish are less reproductively successful than the wild-born descendents of wild fish.  With the completion of the third generation of the pedigree we analyzed the fitness of wild-born fish as a function of their parentage (Fig. 3).  We found that wild-born offspring of two first-generation hatchery fish averaged 37% the fitness of the offspring of two wild fish, while offspring of hatchery-by-wild crosses averaged 87% (Araki et al., 2009).  These results suggest that the hatchery genetic load is not purged from the wild-born population, despite a full generation of natural selection in the wild. 

 Fig. 3A Sampling design to estimate fitness of different types of wild-born fish. Wild-born fish in the mid-1990’s were of three types, depending on whether their parents were both wild, both first-generation hatchery (Hnew), or one of each. We estimated the fitness of these wild-born fish via the return of their adult offspring in the third generation. This figure illustrates the design for estimation of the RRS of fish from run year 99-00 only.  The analysis was carried for three run years (99-00 through 01-02)

Fig 3 again

Figure 3B.  Relative reproductive success of different types of wild-born fish. Fitness of different types of fish relative to that of the ‘best’ type (fish that had two wild parents).  Graphs on left are for fish for which we know both parents.  Graphs on right are for fish for which we have one missing parent (- means missing, presumably a wild resident). Top graph is male breeders, bottom graph is for females.  In this figure we used ‘C’ for captive-bred’, rather than H. 

image005

To summarize the main work to date on the Hood River, we have shown: (1) the older, multi-generation, summer and winter hatchery stocks from the Hood River had very low fitness relative to wild fish (10-30%).  This result is consistent with results of many other studies on old stocks (Berejikian and Ford, 2004; Araki et al., 2008; Recovery Implementation Science Team, 2009). (2) first generation winter run fish have significantly lower fitness than wild fish (about 85%), second generation fish do even worse, and the effect is genetically based. (3) The genetic effects of hatchery culture identified for the winter-run stock persist into the first wild-born generation, with the fitness of wild-born fish depending on whether their parents were both wild, both hatchery or one of each.  Again, the common environment experienced by these three types of wild fish suggests a genetic effect.

 

II. Other questions addressed using the Hood River steelhead system

 (1) Effective population size (Araki et al., 2007c)

We used our long-term pedigree to conduct an extensive analysis of the effects of hatchery fish and resident fish on the effective number of breeders per year (Nb), and on the effective sizes per generation (Ne) in the Hood River.  Using demographic data and genetic parentage analysis on an almost complete sample of all the adults that returned to the river over 15 years (>15,000 individuals), we estimated Nb in 13 run years (5 summer run years and 8 winter run years) and Ne for three entire generations.

Here we found that the ratio of Ne to the estimated whole population size, N, was 0.17-0.40, with large variance in reproductive success among individuals being the primary cause of reduction in Ne/N.  Fish from the traditional hatchery programs had negative effects on Nb, not only by reducing mean reproductive success, but also by increasing variance in reproductive success among breeding parents.  No sign of such effects was found in fish from supplementation hatchery programs.  We observed relatively stable Nb over years and high Ne/N per generation, even though run sizes fluctuated widely during those years.  Also, there is a negative correlation between the anadromous run size and the fraction of missing parents (Fig. 4).  These two observations suggest that a constant-sized pool of resident trout interbreeding with the anadromous fish dampens fluctuations in yearly Nb, and thus enhances long-term Ne.

  

Figure 4. Negative correlation between fraction of parents missing and run size (left panel: summer run; right panel: winter run) suggests that a constant-sized trout population interbreeds with a fluctuating anadromous population.

image006image007

 

(2) Methodological contributions:

            a) Estimation of effective size in supplemented populations (Araki et al., 2007d):

We also used the dataset to evaluate the performance of two indirect methods that are often used to estimate the effective size of populations.  We compared estimates using the temporal method (TEMP) and the linkage disequilibrium (LD) method with the direct estimate available via our pedigree. While the LD estimate was generally in close agreement with the direct estimate, the TEMP estimate was much lower in sample sets that were dominated by nonlocal hatchery fish with low reproductive success.  This bias in the temporal method, which arises when genes associated with a particular group of parents are selected against in the offspring sample, has not been appreciated before. Such situations may be particularly common when artificial propagation or translocations are used for conservation.  Thus, we have identified a useful methodological issue that will be of use in predicting the loss of genetic diversity in supplemented populations.

             b) How to obtain an unbiased estimator of RRS (Araki and Blouin, 2005):

We investigated the effects of two types of assignment error on the bias in an estimate of RRS between two groups (as with hatchery vs. wild).  We derive equations to adjust raw RRS estimates for these assignment errors to obtain unbiased estimates of RRS as well as of the number of offspring whose parents were not sampled.  These methods are now in use by other RRS studies.

 

 (3) Theoretical analysis of the causes of fitness decline (Araki et al., 2008): 

This manuscript has two parts: a review of data on the fitness of hatchery versus wild salmonids with discussion of possible causes, and a quantitative genetic analysis of whether strong selection alone could explain the steep fitness drop we observed in Hood River winter-run fish after a single generation of hatchery culture.  We conclude that selection alone could cause the difference, but that multiple traits must be under selection. 

 

 (4) Fitness of repeat spawners:

A few percentage of steelhead outmigrate after spawning and return in a second or sometimes a third year to attempt to spawn again (iteroparity) (Keefer et al., 2008).  Spent individuals that are returning to sea are known as kelts.  Most kelts and repeat spawners in the Columbia basin are female (Keefer et al., 2008; Evans et al., 2008).  This observation suggests that the lifetime fitness payoff for iteroparity is higher for females than for males.  Fitness includes a component of survival and a component of reproductive success, given they make it back.  We know that female kelts survive better and are more likely to make it back than male kelts (Keefer et al., 2008; Evans et al., 2008).  However, to our knowledge, the hypothesis that the fitness payoff for females that make it back is higher than that for males that make it back has never been tested. 

We now have estimates of lifetime fitness for a large enough sample of repeat spawners from the Hood River to test that hypothesis.  The fitness consequences of spawning more than once do indeed differ substantially between sexes.  Females obtain a huge increase in lifetime fitness by being iteroparous.  Returning females produced as many surviving adult offspring during their second spawning as during their first.  Thus, females essentially double their lifetime fitness by returning a second time.  The situation is very different for males.  Males that returned a second time were those that had low average fitness during their first attempt.  Their second spawning brought their total fitness up to that of an average male that spawned just once.  Thus, it appears that not only do females survive better as kelts than males, but they gain a higher reproductive payoff if they make it back.  We are currently adding one more run year of data to this analysis and plan to write it up for publication in the coming year.

 

 (5) Resident fish – the unsampled parents:

I also describe these results in the section on “Response to past ISRP and Council comments and Recommendations” because this work was initiated in direct response to the ISRP’s 2006 request.  In brief, despite a dam that stops 100% of the anadromous fish, we are missing around 50% of the parents each year, and we have estimated that only a small fraction of those missing parents could be owing to parentage matching errors, fish from other runs, and so on.  Other groups doing pedigree work on steelhead find exactly the same phenomenon; e.g. Seamons et al., 2004).  Thus, there appears to be a resident population that interbreeds freely with the steelhead population.  This is probably the rule rather than the exception for steelhead.  So what effect does that resident component of the population have on the genetics of the steelhead component? And what about residualized hatchery fish, which we know to be common in other river systems?

We previously showed that a constant sized population of residents appears to keep the effective population size of the anadromous population relatively constant despite wild fluctuations from year to year in the anadromous census size (see Figure 2 in Araki et al., 2007c, also Fig. 4 above). Thus, the resident fish appear to be very important for maintaining genetic diversity in the anadromous portion of the population.

Most recently, we developed an indirect method to estimate the fraction of those missing parents that were residualized hatchery fish (Christie et al., in review).  We found that fewer than 10% of the missing parents could have been residualized hatchery fish.  Furthermore, if we consider all the sources of genes that go into producing wild-born steelhead (parents that are wild anadromous, hatchery anadromous, wild residents or residualized hatchery fish), only 1% could be attributed to residualized hatchery fish (Fig. 5).  Thus, residualized hatchery fish are not a significant alternate route for hatchery genes to flow into the wild steelhead population.

Thus, we have identified two ways in which the resident population benefits the wild population: (1) maintaining genetic diversity by buffering against fluctuations in the effective size, and (2) reducing the rate of gene flow from the hatchery into the wild population.

 Figure 5.   Estimated sources of gene flow into anadromous Hood River steelhead. The orange and blue pie slices represent the amount of hatchery and wild gene flow, respectively, into the steelhead ‘population’ from all matings involving anadromous fish (i.e., anadromous x anadromous and anadromous x resident).  Here we account for the lower fitness of anadromous hatchery fish. The red and green slices represent the amount of gene flow into the steelhead ‘population’ from residualized hatchery steelhead (red) and from wild residents (green) via matings between resident and anadromous O. mykiss (estimated from the grandparentage study).  The purple slice represents the amount of gene flow from resident x resident matings, for which we could not estimate the proportion that came from the hatchery. Note that at least 20% of steelhead genes (probably closer to 46%) come from wild resident O. mykiss.

image009

 

 

 

(6) Studies on the mechanisms causing hatchery fish to rapidly decline in fitness

  6a) Possible epigenetic effects of the hatchery (Blouin et al., 2010):

Here we tested whether returning winter-run adults of wild or Hnew origin differed in overall levels of genomic methylation.  They do not appear to.

            

6b) The growth rate hypothesis:

Steelhead are usually raised quickly and released at one year of age, even though they normally spend two years in freshwater.  Size at smolting is correlated with survival to return (Reisenbichler et al., 2004). In the high-food and predator-free hatchery environment, this survival difference should select for high growth rate in hatchery juveniles, perhaps via behavioral or metabolic changes.  But an excessively high growth rate can be maladaptive in natural environments (Arendt, 1997).  So when those surviving hatchery fish reproduce in a natural setting, their offspring could have lower survival than wild offspring owing to inappropriate behaviors, an excessively high metabolism, or other maladaptive consequences of a genetic predisposition for fast growth in the benign and high-food hatchery environment (Sundstrom et al., 2005; Biro et al., 2004; Tymchuk et al., 2007).  Furthermore, it is a well established paradigm of life history theory that populations evolve to optimize unavoidable tradeoffs among growth, reproduction and maintenance (e.g. Blair and Wolfe, 2004).  Thus, selection for growth rate is a very compelling explanation because it could influence both reproductive performance of the hatchery adults themselves and the survival of any offspring they do produce.

 Therefore, we have begun testing whether there is evidence that the Hood River hatchery fish have been selected to have higher growth rates under hatchery conditions.  In 2008 we set up a series of 2x2 matrix crosses involving a hatchery and wild male by a hatchery and wild female, and then raised each pooled group of four families in the hatchery for seven weeks.  We measured the size of each fish and archived a tissue sample for DNA typing.  After sorting them back into their respective families, we found no evidence for an additive genetic effect of hatchery versus wild background, although there was a weak maternal effect in which the offspring of wild females seemed to grow faster (Fig. 6) (perhaps an effect of egg size? e.g. Heath et al., 2003).  We started a second set of crosses in 2009 and sampled them at 2 months of age and one year of age. Among the 2-month olds we again found little evidence for a difference between fish of H vs. W backgrounds.  We now have their siblings raised to one year of age, and plan to analyze them the same way.  It is possible that any growth difference will appear later in life, so we haven’t ruled out the growth rate hypothesis yet.  Finally, we are setting up a 3rd set of crosses this summer, and will replicate the experiment in one final year, this time raising them at higher densities to more closely match those of the production fish (last year’s fish were not as crowded).  If we still see no effect of genetic background, then either the growth rate hypothesis is not true, or the growth rate difference is expressed under very particular environmental conditions that we failed to replicate.

 Figure 6.  2008 growth rate experiment.  Here we show, plotted on the same graph, the sizes of all individuals from 15 2x2 matrices, where each matrix was created by crossing a hatchery and wild male with a hatchery and wild female (approximately 15 offspring per full sib family).  There is no significant main effect of type of father within females, as expected if there is an additive genetic effect of fish type.  However, there were strong maternal effect within each matrix, and on average, the offspring of wild mothers grew faster (WxW and WxH columns in the figure).

image010

 

One intriguing result of the 2009 growth rate experiments was that although there was no significant effect of fish type, there was a significant effect of cross date – families spawned earlier grew faster (Fig. 7).  Each year the Hood River production crosses are created over about a seven week period.  The earlier families are chilled until the last ones are done, then they all develop together and are ponded simultaneously.  They all experience the same total thermal units, but it is possible that the slightly different environmental treatments cause size variation among families.  Such an environmental effect could have survival consequences later in life, thus inflating variance in family numbers.  Alternately, there may be some subtle selection going on in the choice of broodstock (cf. McLean et al., 2005).  Perhaps broodstock that are chosen first (or those that are ready the earliest) produce faster growing offspring.  We are currently trying to tease apart these possibilities (for example, cross date is correlated with run timing of the mothers, but run timing explains less of the variance in family size than does cross date).

Figure 7. Effect of cross date on family mean growth rate.  Here we had 14 HxH familes (left) and 18 WxW families (right).  Crosses made earlier in the season grew faster, even though all families were started at the same sizes and were ponded together on the same date.  In this experiment all fish were raised in the same tank and were sorted back into families after sampling.

image011

  

6c) Metabolic rate

We also tested the hypothesis that fish of HxH and WxW background differ in resting metabolic rate.  This hypothesis was inspired by the observation that older hatchery stocks at the Oak Springs hatchery appear to feed continuously over the winter whereas the first-generation stocks substantially reduce their feeding rate during that time (Lyle Curtis, hatchery manager, pers. comm.). One mechanism by which a population could respond to selection for large size in one year would be to maintain an elevated metabolic rate and feeding rate during the winter months when fish in the wild normally become quiescent.  Thus, the hatchery may be selecting for a “feed-all-winter morph”, a phenotype that presumably has costs when expressed in the wild. Therefore, in mid-winter (January 2010) we measured the resting metabolic rate of 15 juveniles of HxH and 15 of WxW background, raised in the same conditions. We found no evidence of a difference (Fig. 8).  So, if there is a growth rate difference, we can probably rule out metabolic rate as the proximal mechanism.

 Figure 8Metabolic rate experiment. Total oxygen consumption per day, measured individually on each of 15 fish of HxH ancestry and 15 fish of WxW ancestry (siblings of the 2009 growth rate families). There is no effect of fish type.

image012

 

  6d) Disease resistance

As a corollary to the hypothesis that hatchery fish have been selected for fast growth, we are also testing whether they have reduced disease resistance.  Again, there are unavoidable tradeoffs among growth, reproduction and maintenance (which includes disease resistance).  Thus, selection for increased juvenile growth rate could very easily result in a tradeoff with maintenance functions.  Furthermore, hatchery fish are raised in a relatively disease-free environment, so one or more generations of relaxed natural selection could also contribute to their low fitness when exposed to pathogens in the wild.  In collaboration with Jeri Bartholomew (OSU Dept. Microbiology), we are currently testing whether fish of HxH, HxW, WxH and WxW background differ in their susceptibility to bacterial and viral pathogens (Vibrio, IHNV). 

 

  6e) Possible sex reversal or gender disruption caused by the hatchery environment?

Interestingly, the sex ratio of hatchery fish at Powerdale dam is less female biased than the sex ratio of returning wild fish.  Gender expression in fish can be very labile, especially in response to temperature (Craig et al., 1996; Conover, 2004).  The warmer water of the hatchery would be consistent with an excess of males, as we observed relative to wild fish (Ospina-Alvarez and Piferrer, 2008).  Therefore, we used a PCR-based sexing method (Brunelli et al., 2008) to determine the chromosomal sex of several hundred wild and hatchery fish, of both summer and winter runs.  In short, the genders identified by the ODFW staff at the dam matched the chromosomal sexes very well, with no trend towards excess mis-identifications in any type of fish.  Another beautiful hypothesis slain by ugly facts.

 Nevertheless, an interesting side result of this study was that we showed that the ODFW staff are very good at sexing returning summer run fish.  This can be very difficult owing to the fact that summer run fish show up relatively immature (sometimes up to almost year before they breed).  The sex ratio difference between returning hatchery and wild fish remains an interesting, unexplained phenomenon. We are currently exploring the hypothesis that it results from different patterns of sex-specific survival.

 

 

I hope the summary of our work above shows that we have been productive with the funding we have been granted to date.  Below is a list of our publications from the Hood River.  We will soon be writing up many of the unpublished results described above (repeat spawners, growth rate & metabolic rate, sex change).  We are just beginning the disease resistance work, which we hope will produce publishable results next year.  Those studies will tie into the work on relaxed selection on MHC that we propose for the future.

 

Manuscripts to date that resulted directly from our work on the Hood River(all but the last have been uploaded to PISCES and are available in the “Existing Project Documents” link on the cbfish site)

Araki, H. and M.S. Blouin. 2005. Unbiased estimation of relative reproductive success of different groups: evaluation and correction of bias caused by parentage assignment errors. Molecular Ecology, 13:4907-4110.

 Araki, H., W.R. Ardren, E. Olsen, B. Cooper and M.S. Blouin. 2007a. Reproductive success of captive-bred steelhead trout in the wild: evaluation of three hatchery programs in the Hood River. Conservation Biology, 21:181-190. 

 Araki, H., B. Cooper and M.S. Blouin. 2007b. Genetic effects of captive breeding cause a rapid, cumulative fitness decline in the wild. Science, 318: 100-103.

 Araki, H., R.S. Waples, W.R. Ardren, B. Cooper and M.S. Blouin. 2007c. Effective population size of steelhead trout: influence of variance in reproductive success, hatchery programs, and genetic compensation between life-history forms. Molecular Ecology 16:953-966

 Araki, H., R.S. Waples and M.S. Blouin. 2007d. A potential bias in the temporal method for estimating Ne in admixed populations under natural selection. Molecular Ecology 16: 2261–2271

 Araki, H., B. Berejikian, M. Ford, and M.S. Blouin. 2008 Fitness of hatchery-reared salmonids in the wild.  Evolutionary Applications 1:342-355. 

 Araki, H., B. Cooper and M.S. Blouin. 2009. Carry-over effect of captive breeding reduces reproductive fitness of wild-born descendants in the wild. Biology Letters 5: 621-624.

 Blouin, M.S. V. Thuillier, B. Cooper, V. Amarasinghe, L. Cluzel, H. Araki and C. Grunau. 2010. No evidence for large differences in genomic methylation between wild and hatchery steelhead trout (Oncorhynchus mykiss). Canadian Journal of Fisheries and Aquatic Sciences. 67: 217-224.

 Christie, M., M.L. Marine and M.S. Blouin. Who are the missing parents? Grandparentage analysis identifies multiple sources of gene flow from hatchery fish into a wild population. Molecular Ecology, In review. [draft in review to be provided to the ISRP]

 



The table content is updated frequently and thus contains more recent information than what was in the original proposal reviewed by ISRP and Council.

Review: 2018 Research Project Status Review

Council Recommendation

Assessment Number: 2003-054-00-NPCC-20210302
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: 2018 Research Project Status Review
Approved Date: 12/20/2018
Recommendation: Implement
Comments: This set of projects [200303900, 200305400, 200306300 and 201003300] went through a policy review in 2017, and this review by the ISRP for progress. Studies to date have revealed that RRS between hatchery and naturally spawning fish can be reduced in a variety of ways. Because of this complexity, a more detailed conceptual framework is needed to predict how different species or populations will respond to hatchery supplementation and to allow managers to make better case-specific decisions. The ISRP believes that an updated synthesis is needed to make progress toward such a framework. They suggest that any new effort to synthesize results across the RRS studies should consider the history of hatchery influence prior to and
during each study. Many of the projects reviewed are expected to report their most valuable results over the next few years. At that time, an updated synthesis of findings will be especially valuable. The ISRP is reassured that the RRS studies are on track and that proponents are collaborating and sharing information effectively. The Council concurs and asks that the sponsors work together on a synthesis report to be submitted and reviewed by the Council and the ISRP ahead of the start of the 2021 Anadromous Habitat and Hatchery Review process.

Recommendation: Bonneville to work with the sponsors on a coordinated reporting of results as a “synthesis” review. Bonneville and the sponsors are requested to present this progress report/results to the Council and ISRP in summer of 2020; close to when these projects will be wrapping up, and ahead of the 2020 Anadromous Habitat and Hatchery Review.
Review: RME / AP Category Review

Council Recommendation

Assessment Number: 2003-054-00-NPCC-20110124
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: RME / AP Category Review
Proposal: RMECAT-2003-054-00
Proposal State: Pending BPA Response
Approved Date: 6/10/2011
Recommendation: Fund (Qualified)
Comments: Implement through FY 2014. Implementation beyond 2014 based on ISRP and Council review of the results report and/or outcome of a regional hatchery effects evaluation process.
Conditions:
Council Condition #1 Programmatic Issue: RMECAT #4 Hatchery Effectiveness—.
Council Condition #2 Programmatic Issue: RMECAT #6 Research projects in general—.

Independent Scientific Review Panel Assessment

Assessment Number: 2003-054-00-ISRP-20101015
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: RME / AP Category Review
Proposal Number: RMECAT-2003-054-00
Completed Date: 12/17/2010
Final Round ISRP Date: 12/17/2010
Final Round ISRP Rating: Meets Scientific Review Criteria
Final Round ISRP Comment:
This project was initiated in 2003 as one of a suite of projects to evaluate relative reproductive success of hatchery-origin steelhead compared to natural steelhead when spawning naturally, to address critical uncertainties identified in the 2000 FCRPS BiOp RPA 182.
First Round ISRP Date: 10/18/2010
First Round ISRP Rating: Meets Scientific Review Criteria
First Round ISRP Comment:

This project was initiated in 2003 as one of a suite of projects to evaluate relative reproductive success of hatchery-origin steelhead compared to natural steelhead when spawning naturally, to address critical uncertainties identified in the 2000 FCRPS BiOp RPA 182.

Documentation Links:
Review: FY07-09 Solicitation Review

Council Recommendation

Assessment Number: 2003-054-00-NPCC-20090924
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: FY07-09 Solicitation Review
Approved Date: 10/23/2006
Recommendation: Fund
Comments:

Independent Scientific Review Panel Assessment

Assessment Number: 2003-054-00-ISRP-20060831
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: FY07-09 Solicitation Review
Completed Date: 8/31/2006
Final Round ISRP Date: None
Final Round ISRP Rating: Meets Scientific Review Criteria
Final Round ISRP Comment:
The response addressed the ISRP questions. The ISRP appreciated the effort to address the review in a professional and positive manner with explanatory notes and even figures. The ISRP expects that the principal investigators will consider the ISRP's comments on residualized hatchery fish in subsequent proposals, reports, and reviews.

A thorough response and additional references were provided, for the most part. Clearly, this is important work on the issue of wild and hatchery fish interactions and supplementation. The papers in press, in review, and planned shall become important contributions to fisheries science and particularly to the question of supplementation in the Columbia River Basin. The opportunity to review the papers in press or in review was much appreciated and assisted in confirming or addressing previous ISRP concerns quite adequately.

The question of contribution of residualized hatchery fish to parentage of wild and hatchery returns remains. Htrad may have provided no evidence of a parental contribution to returns since their success in spawning (or of progeny post-spawning) may have been near zero, but Hnew males may be more successful. The implications of reproductive success of residualized Hnew males may be substantial. It seems this could be addressed with more planning and thought, perhaps by sub-sampling residuals directly or by samples from hatchery smolts released at acclimation sites throughout the Hood River. Indeed, the opportunity may be unique to this system. Does "acclimatization" provide a benefit or loss to overall reproductive success of wild fish?

Supplementation was shown here (paper in review) to have no effect on the reproductive success of wild fish. However, does it add anything? In other words, if there is no added benefit when wild fish are seeding habitat to capacity, then what is the point of supplementation?

Ecological effects remain an issue. Regardless, a continuation of this work is highly recommended since it will address important questions on the genetics of salmonids and hatcheries, particularly if more focus is placed on the residual steelhead issue, and success in sampling can continue with the removal of the Powerdale Dam, which seems possible.

Further collaboration should be encouraged - this work should form part of a basinwide study on supplementation, filling gaps not possible in other studies and replicating work elsewhere, thus agreement on standard procedures is necessary, as appears to be unfolding.
Documentation Links:
Explain how your project has responded to the above ISRP and Council qualifications, conditions, or recommendations. This is especially important if your project received a "Qualified" rating from the ISRP in your most recent assessment. Even if your project received favorable ratings from both the ISRP and Council, please respond to any issues they may have raised.
Response to past ISRP and Council comments and recommendations: View instructions
The main suggestion in the ISRP review of 8/31/2006 was that we address the issue of whether hatchery fish released into the Hood River residualize and mate with steelhead. This is a question we have wrestled with for several years because despite sampling almost 100% of the anadromous fish at the dam every year, we are missing many parents. Estimating the genetic contribution by residualized hatchery fish is essential because we now know that even first-generation (Hnew) hatchery fish put a genetic load on the wild population. Thus, residualized hatchery fish could be an important, alternate route for hatchery genes to flow into the wild population. Modeling the overall fitness effects of adding hatchery fish to a river will need to take into account this alternate route.<br/> <br/> Our ODFW colleagues (Rod French and co.) have tried for several years to sample the resident population for us, but the upper reaches of the Hood River are very difficult to work in, and they capture very few trout – even though we know they must be out there. Thus, we can’t use a traditional parentage approach. So this year we came up with an indirect way to estimate the number of residualized hatchery fish that were breeding with steelhead (made possible now that we have an almost four generation pedigree). In short, we took the returned fish for whom we could find only one parent, and then tested whether the parents of the missing parent were a known pair of broodstock. We developed some new statistical methods for grandparent-grandoffspring matching and estimation of the probabilities of match by chance. We found that less than 10% of the missing parents could have been residualized hatchery fish. We further estimate that if we consider all the sources of genes that go into producing wild-born steelhead (parents that are wild anadromous, hatchery anadromous, wild trout or residualized hatchery fish), only 1% could be attributed to residualized hatchery fish. Thus, residualized Hnew fish are not contributing substantially to the wild-born anadromous gene pool. Whether this means that few of the Hnew hatchery fish residualize, or that they do but rarely mate with steelhead remains to be seen. However, screw trap data from tributaries and the main stem of the Hood River show that 99% of the Hnew fish leave the river within a month of release, and they are rarely found thereafter; R. French pers. comm..; see also Kostow 2004). Thus, the simplest explanation is that Hnew fish do not residualize in great numbers in the Hood River. The manuscript describing this work is in review for publication. I will provided a draft copy for the ISRP to review. <br/> <br/> An interesting consequence of the above result is that the resident wild trout population is acting as a buffer to reduce the rate of flow of hatchery genes into the wild population of steelhead. This is the second way we have identified in which the resident population is beneficial to the anadromous population. Previously, we presented evidence that a constant-sized population of resident trout is keeping the effective size of the anadromous component relatively constant, even though the anadromous run sizes fluctuate wildly from year to year (Araki et al., 2007c). Thus, the resident component of the population appears essential for both maintaining genetic diversity in the population as a whole, and for mitigating the effects of gene flow from the hatchery into the wild population. <br/> <br/> In the 4th paragraph of the review, the ISRP asked whether supplementation is actually helping the wild population, given our initial data showed no obvious harm. Having now more than doubled the number of years of data since that ISRP review, we now know: <br/> (1) that Hnew fish do indeed have significantly lower fitness than wild (Araki et al., 2007b). <br/> (2) that the wild-born offspring of Hnew fish also suffer a fitness deficit (Araki et al., 2009). <br/> Thus the hatchery fish probably are a negative impact on the wild population.<br/> <br/> In the last paragraph, the ISRP encourages more collaboration with other projects. We and the other groups that are conducting relative fitness studies regularly share our latest results informally and at meetings. Recent reviews of the topic attest to the growing pool of information we now all draw upon (e.g. Recovery Implementation Science Team, 2009; Araki et al., 2008).


Project Level: Please discuss how you’ve changed your project (objectives, actions, etc) based on biological responses or information gained from project actions; because of management decisions at the subbasin state, regional, or agency level; or by external or larger environment factors. Specifically, regarding project modifications summarize how previous hypotheses and methods are changed or improved in this updated proposal. This would include project modifications based on information from recent research and literature. How is your new work different than previous work, and why?
Management Level: Please describe any management changes planned or made because of biological responses or information gained from project actions. This would include management decisions at the subbasin, state, or regional level influenced by project results.
Management Changes: View instructions
There is little evidence that supplementation helps wild populations (e.g. Waples et al., 2007), and growing evidence that it might be harmful. Our data were the first from what is now a growing number of studies showing that even first-generation hatchery fish have lower fitness than wild fish. Data such as these have spurred a re-evaluation of the wisdom of supplementation programs. Managers are now also reconsidering whether integrated broodstock programs (using wild fish as broodstock) are really the panacea we thought they’d be for hatcheries whose mission is harvest. This is because although first-generation hatchery fish are fitter than multi-generation fish, they usually mix much more extensively with wild populations (e.g. by having identical run timings). Thus, completely segregated programs (i.e. going back to using domesticated stocks for harvest but keeping them out of wild spawning grounds, such as via weirs) might actually make more sense in some situations. In one concrete example, based on our work, ODFW is considering such a shift from an integrated to a segregated program on the Hood River (Charlie Corrarino, ODFW Conservation and Recovery Program Manager, pers. comm.).

The table content is updated frequently and thus contains more recent information than what was in the original proposal reviewed by ISRP and Council.

Public Attachments in CBFish

ID Title Type Period Contract Uploaded
00015883-1 Reproductive Success – Steelhead in the Hood River Progress (Annual) Report 11/2003 - 09/2004 15883 10/1/2004 12:00:00 AM
00019502-1 Reproductive Success – Steelhead in the Hood River Progress (Annual) Report 10/2004 - 09/2005 19502 10/1/2005 12:00:00 AM
00019502-2 Reproductive Success – Steelhead in the Hood River Progress (Annual) Report 10/2005 - 09/2006 19502 9/1/2006 12:00:00 AM
P104253 fy07 annual report Progress (Annual) Report 10/2006 - 09/2007 29562 10/26/2007 4:10:53 PM
P108613 Reproductive Success -- Steelhead in the Hood River Progress (Annual) Report 10/2007 - 09/2008 35294 10/13/2008 2:58:43 PM
P109816 Fitness of hatchery-reared salmonids in the wild Other - 39163 1/15/2009 10:49:17 AM
P112132 Hood River publication Other - 39163 6/19/2009 10:20:40 AM
P113759 Hood River steelhead fitness 2008-2009 final report Progress (Annual) Report 07/2009 - 09/2009 39163 10/14/2009 1:05:28 PM
P117053 Araki & Blouin 2005 MOLEC ECOL.pdf Other - 7/9/2010 4:00:20 PM
P117054 Araki et al 2007a CONSERV BIOL.pdf Other - 7/9/2010 4:03:00 PM
P117055 Araki et al 2007b SCIENCE.pdf Other - 7/9/2010 4:04:54 PM
P117056 Araki et al 2007c MOLEC ECOL.pdf Other - 7/9/2010 4:06:22 PM
P117057 Araki et al 2007d MOLEC ECOL.pdf Other - 7/9/2010 4:08:07 PM
P117058 Araki et al 2009 BIOLOGY LETTERS.pdf Other - 7/9/2010 4:09:47 PM
P117059 Araki et al 2008 EVOL APPLIC.pdf Other - 7/9/2010 4:11:27 PM
P117060 Blouin et al 2010 CJFAS.pdf Other - 7/9/2010 4:13:27 PM
P117413 Letter of collaboration for RMECAT proposal - Mockler Other - 7/29/2010 12:46:54 PM
P117414 Letter of collaboration for RMECAT proposal - Jacobson Other - 7/29/2010 12:48:16 PM
P118537 Reproductive Success -- Steelhead in the Hood River, 2009 - 2010 Progress (Annual) Report 10/2009 - 09/2010 43555 10/27/2010 12:01:26 PM
P120891 Christie et al 2011 MOLEC ECOL.pdf Other - 4/17/2011 12:53:48 PM
P123732 Reproductive Success - Steelhead in the Hood River Progress (Annual) Report 10/2010 - 09/2011 49886 11/10/2011 2:42:01 PM
P128503 Christie et al 2012 PNAS.pdf Other - 54599 10/8/2012 3:42:43 PM
P128505 Christie et al_ 2012_ HEREDITY.pdf Other - 54599 10/8/2012 3:43:48 PM
P128554 Reproductive Success - Steelhead in the Hood River; 10/11 - 9/12 Progress (Annual) Report 10/2011 - 09/2012 54599 10/11/2012 12:42:27 PM
P135610 Christie et al 2013 J HEREDITY Other - 58865 1/17/2014 9:50:04 AM
P135611 Christie et al. 2013 BIOINFORMATICS Other - 58865 1/17/2014 9:51:57 AM
P135613 Christie 2013 MOLECULAR ECOLOGY Other - 58865 1/17/2014 10:07:13 AM
P138528 Reproductive Success - Steelhead in the Hood River; 1/13 - 12/13 Progress (Annual) Report 01/2013 - 12/2013 63427 8/28/2014 2:00:13 PM
P141899 Christie et al 2014 EVOLUTIONARY APPLICATIONS Other - 1/23/2015 2:37:19 PM
P141901 Fox et al. 2014 MARINE GENOMICS Other - 1/23/2015 2:42:37 PM
P141902 Thompson et al. 2014. Environmental Biology of Fishes Other - 1/23/2015 2:45:25 PM
P143542 Reproductive Success - Steelhead in the Hood River; 1/14 - 12/14 Progress (Annual) Report 01/2014 - 12/2014 67127 6/1/2015 2:14:17 PM
P148412 Christie et al. 2016 NATURE COMMUNICATIONS.pdf Other - 3/4/2016 2:24:21 PM
P148413 Thompson and Blouin 2015 CANADIAN J FISHERIES & AQUATIC SCI Other - 3/4/2016 2:27:43 PM
P148409 Reproductive Success - Steelhead in the Hood River; 1/15 - 12/15 Progress (Annual) Report 01/2015 - 12/2015 70177 4/14/2016 9:13:16 AM
P154341 Thompson et al.2016 Environmental Biology of Fishes Other - 70177 2/24/2017 12:37:11 PM
P154342 Thompson and Blouin 2016 Trans Am Fish Soc Other - 70177 2/24/2017 12:44:12 PM
P154373 Reproductive Success -- Steelhead in the Hood River; 1/16 - 12/16 Progress (Annual) Report 01/2016 - 12/2016 74233 4/17/2017 1:45:00 PM
P156692 Bibliography to 2016 Other - 8/3/2017 11:53:51 AM
P159756 Reproductive Success - Steelhead in the Hood River; 1/17 - 12/17 Progress (Annual) Report 01/2017 - 12/2017 77259 3/16/2018 4:24:28 PM
P159780 Thompson et al. 2018 Aquaculture Other - 77259 3/20/2018 9:19:28 AM
P160357 Christie. et al 2018 PNAS life history Other - 77259 5/3/2018 3:04:23 PM
P164363 Reproductive Success - Steelhead in the Hood River; 1/18 - 12/18 Progress (Annual) Report 01/2018 - 12/2018 76914 REL 2 3/11/2019 12:06:14 PM
P176835 Reproductive Success - Steelhead in the Hood River; 1/19 - 12/19 Progress (Annual) Report 01/2019 - 12/2019 76914 REL 4 6/23/2020 5:24:09 PM

Other Project Documents on the Web

None


The Project Relationships tracked automatically in CBFish provide a history of how work and budgets move between projects. The terms "Merged" and "Split" describe the transfer of some or all of the Work and budgets from one or more source projects to one or more target projects. For example, some of one project's budget may be split from it and merged into a different project. Project relationships change for a variety of reasons including the creation of efficiency gains.
Project Relationships: None

Additional Relationships Explanation:

A. Geographic Region

This project is coordinated with the Hood River steelhead hatchery and research program, funded by BPA and administered and implemented by the ODFW and Warm Springs Tribes under the projects listed below.  

198805303 Hood River Production Monitoring and Evaluation (M&E) - Warm Springs
198805304 Hood River Production Monitor and Evaluation (M&E) - ODFW
198805307 Hood River Production Operations and Maintenance (L&M) - Warm Springs
198805308 Hood River Production Operations and Maintenance (L&M) – ODFW

Note that we have also been collaborating with the Warm Springs tribe’s Parkdale hatchery and the ODFW’s Oak Springs hatchery to produce and raise individuals of wild,  hatchery or mixed background in order to test whether individuals of different genetic backgrounds, but raised in a common garden environment, differ in traits such as growth rate, metabolic rate, propensity to smolt, disease resistance and whole-genome patterns of gene expression.  

B. Similar Work:
There are a number of studies ongoing that have components involving estimating the relative reproductive success of fish of various types.  Each has a unique design and set of peculiar circumstances.  However, we share our results regularly at meetings, and eventually it will be possible to do a meta-analysis of results from studies on different species, life history types, geographic regions, and so on, in order to elucidate any generalities about the fitness of hatchery and wild fish.  BPA-funded studies having a RRS component include the following:

Genetic Monitoring and Evaluation (M&E) Program for Salmon and Steelhead 198909600

Grand Ronde Early Life History of Spring Chinook and Steelhead 199202604

Grande Ronde Supplementation Operations and Maintenance (O&M) and Monitoring and Evaluation (M&E) on Lostine River 199800702

Grande Ronde Supplementation O&M on Catherine Creek/Upper Grande Ronde River 199800703

Grande Ronde Spring Chinook on Lostine/Catherine Creek/ Upper Grande Ronde Rivers
199800704

Monitor and Evaluate (M&E) Reproductive Success and Survival in Wenatchee River 200303900

Evaluate the Relative Reproductive Success of Wild and Hatchery Origin Snake River Fall Chinook Spawners Upstream of Lower Granite Dam 200306000

Grande Ronde Supplementation Monitoring and Evaluation (M&E) on Catherine Creek/Upper Grande Ronde River 200708300

Study Reproductive Success of Hatchery and Natural Origin Steelhead in the Methow 201003300

Evaluate the Reproductive Success of Wild and Hatchery Steelhead in Natural and Hatchery Environments  2003-050-00

Evaluate Reproductive Success of Kelt Steelhead  2003-062-00

Natural Reproductive Success and Demographic Effects of Hatchery-Origin Steelhead in Abernathy Creek, Washington  2003-063-00

Investigation of Relative Reproductive Success of Stray Hatchery & Wild Steelhead & Influence of Hatchery Strays on Natural Productivity in Deschutes  2007-299-00

Genetic Monitoring and Evaluation (M&E) Program for Salmon and Steelhead 1989-096-00


Primary Focal Species
Steelhead (O. mykiss) - Lower Columbia River DPS (Threatened)

Secondary Focal Species
None

Describe how you are taking into account potential biological and physical effects of factors such as non-native species, predation increases, climate change and toxics that may impact the project’s focal species and their habitat, potentially reducing the success of the project. For example: Does modeling exist that predicts regional climate change impacts to your particular geographic area? If so, please summarize the results of any predictive modeling for your area and describe how you take that into consideration.
Threats to program investments and project success: View instructions
NA

Work Classes
Please explain how you manage the data and corresponding metadata you collect.
<No answer provided>
Describe how you distribute your project's data to data users and what requirements or restrictions there may be for data access.
<No answer provided>
What type(s) of RM&E will you be doing?
Uncertainties Research (Validation Monitoring and Innovation Research)
Where will you post or publish the data your project generates?

Loading ...
Layers
Legend
Name (Identifier) Area Type Source for Limiting Factor Information
Type of Location Count
Hood River (1707010507) HUC 5 EDT (Ecosystem Diagnosis and Treatment) 22

Project Deliverable definition: A significant output of a project that often spans multiple years and therefore may be accomplished by multiple contracts and multiple work elements. Contract Deliverables on the other hand are smaller in scope and correspond with an individual work element. Title and describe each Project Deliverable including an estimated budget, start year and end year. Title: A synopsis of the deliverable. For example: Crooked River Barrier and Channel Modification. Deliverable Description: Describe the work required to produce this deliverable in 5000 characters or less. A habitat restoration deliverable will contain a suite of actions to address particular Limiting Factors over time for a specified Geographic area typically not to exceed a species population’s range. Briefly include the methods for implementation, in particular any novel methods you propose to use, including an assessment of factors that may limit success. Do not go into great detail on RM&E Metrics, Indicators, and Methods if you are collecting or analyzing data – later in this proposal you’ll be asked for these details.
Project Deliverables: View instructions
Estimate RRS of summer run fish as function of ancestry (DELV-1)
1. Estimate RRS of summer run Hnew vs. wild fish in each of 4 run years.

2. Summarize fitness effects of ancestry based on entire 19-year pedigree (here ancestry includes parents that were wild for at least one generation, Skamania hatchery stock, Hnew stock, or unsampled resident trout).

3. Estimate the effect of hatchery ancestry (Skamania stock) on the performance of broodstock in the conservation hatchery. Here the hypothesis is that wild fish used as broodstock will perform better in the hatchery (produce more returning hatchery offspring) if their ancestry contains a large fraction of hatchery genes. The corollary hypothesis is their successfully returning offspring will have poorer RRS in the wild.
Types of Work:
Work Class Work Elements
Research, Monitoring, and Evaluation + Data Management
157. Collect/Generate/Validate Field and Lab Data
162. Analyze/Interpret Data
183. Produce Journal Article

Estimate RRS of winter-run fish (DELV-2)
1. Refine point estimates of RRS of Hnew fish published previously by adding 3 more run years of data.

2. Summarize fitness effects of ancestry based on entire 19-year pedigree (here ancestry includes parents that were wild for at least one generation, Hnew hatchery stock, or unsampled resident trout).

3. Estimate the effect of hatchery ancestry (Hnew fish) on the performance of wild broodstock used in the conservation hatchery. The hypothesis is that wild fish used as broodstock will perform better in the hatchery (produce more returning hatchery offspring) if their ancestry contains a large fraction of hatchery genes, but their offspring will perform more poorly.
Types of Work:
Work Class Work Elements
Research, Monitoring, and Evaluation + Data Management
157. Collect/Generate/Validate Field and Lab Data
162. Analyze/Interpret Data
183. Produce Journal Article

Analysis of differential gene expression between fry of different ancestry (DELV-3)
See "Problem Statement/technical background" for more details, including methods.

Screen the genomes of fish of different ancestry (HxH, HxW, WxH and WxW) that were raised in a common garden to identify loci that are differentially expressed. Deliverable will be a list of differentially expressed loci, an analysis of the physiological pathways they control, and a list of hypotheses about the traits that might be under selection.
Types of Work:
Work Class Work Elements
Research, Monitoring, and Evaluation + Data Management
157. Collect/Generate/Validate Field and Lab Data
162. Analyze/Interpret Data
183. Produce Journal Article

Estimate effects of relaxed selection on immunity genes (MHC) (DELV-4)
See "Project Statement/Technical Background" for further details, including methods.

Here we will conduct the following:

1. Test whether MHC genotype of wild and hatchery fish from the Hood River is correlated with fitness, including
lifetime reproductive success (as estimated from the pedigree),
disease resistance in disease challenge trials, and
parasite load in returning adult fish.

2. Test whether random choice and pairing of broodstock in the hatchery leads, on average, to less desirable genotypes than found in wild fish.

3. Test whether broodstock having compatible MHC genotypes produced more surviving offspring, and surviving offspring with higher RRS

4. Estimate fitness benefit that would be achieved by pairing broodstock based on their MHC, rather than randomly.
Types of Work:
Work Class Work Elements
Research, Monitoring, and Evaluation + Data Management
157. Collect/Generate/Validate Field and Lab Data
162. Analyze/Interpret Data
183. Produce Journal Article


Objective: Describe fitness effects of hatchery summer run steelhead (OBJ-1)

Project Deliverables How the project deliverables help meet this objective*

Estimate RRS of summer run fish as function of ancestry (DELV-1)


Objective: Describe fitness effects of hatchery winter-run steelhead (OBJ-2)

Project Deliverables How the project deliverables help meet this objective*

Estimate RRS of winter-run fish (DELV-2)


Objective: Identify genes differentially expressed between fry of H and W ancestry (OBJ-3)

Project Deliverables How the project deliverables help meet this objective*

Analysis of differential gene expression between fry of different ancestry (DELV-3)


Objective: Test effects of relaxed selection on immunity genes (MHC) (OBJ-4)

Project Deliverables How the project deliverables help meet this objective*

Estimate effects of relaxed selection on immunity genes (MHC) (DELV-4)


*This section was not available on proposals submitted prior to 9/1/2011

Project Deliverable Start End Budget
Estimate RRS of summer run fish as function of ancestry (DELV-1) 2012 2013 $320,000
Estimate RRS of winter-run fish (DELV-2) 2013 2015 $320,000
Analysis of differential gene expression between fry of different ancestry (DELV-3) 2012 2014 $320,000
Estimate effects of relaxed selection on immunity genes (MHC) (DELV-4) 2014 2015 $319,522
Total $1,279,522
Requested Budget by Fiscal Year

Fiscal Year Proposal Budget Limit Actual Request Explanation of amount above FY2010
2012 $305,840 5% increase over FY 2011. This is necessary to keep up with our ability to do the work after four years of flat budgets from FY 2006-2009 (e.g. personnel benefits costs are now substantially higher). Detailed budget justification available upon request. See also timeline in “Problem Statement&quot;.
2013 $315,016 FY 2012 + 3% inflation
2014 $324,466 FY 2013 + 3% inflation
2015 $334,200 FY 2014 + 3% inflation
Total $0 $1,279,522
Item Notes FY 2012 FY 2013 FY 2014 FY 2015
Personnel $169,782 $174,875 $180,121 $185,525
Travel $0 $0 $0 $0
Prof. Meetings & Training $2,000 $2,060 $2,121 $2,185
Vehicles $1,500 $1,545 $1,591 $1,639
Facilities/Equipment (See explanation below) $0 $0 $0 $0
Rent/Utilities $0 $0 $0 $0
Capital Equipment $0 $0 $0 $0
Overhead/Indirect $93,558 $96,366 $99,258 $102,235
Other lab supplies & services $39,000 $40,170 $41,375 $42,616
PIT Tags $0 $0 $0 $0
Total $305,840 $315,016 $324,466 $334,200
Major Facilities and Equipment explanation:
Blouin’s lab occupies 1700 sq. ft. of laboratory space in Cordley Hall at Oregon State University. This includes wet labs, separate clean-rooms for PCR set-up, a conference area, and computers. All postdocs and students have office space. We have all the equipment necessary to conduct the molecular experiments described in this proposal, including multiple thermocyclers and a Robbins Scientific Hydra-96 liquid handling robot. Blouin is a member of the Oregon State University Center for Genome Research and Biocomputing (CGRB; http://www.cgrb.orst.edu/). The CGRB is located in an adjoining building and provides various services such as sequencing and microsatellite genotyping on an ABI 3730 capillary sequencer, and equipment to support genomics and biocomputing research on campus. This includes the Illumina 1G Genome Analyzer that is at the core of the gene expression proposal (http://www.illumina.com/pages.ilmn?ID=203). Computational resources available to this project include the Mockler lab’s dedicated 28-node Linux cluster (Brachy), web and database servers (OCP. Skynet, MCP). In addition, the OSU Center for Genome Research and Biocomputing (CGRB; http://www.cgrb.orst.edu/) maintains an active biocomputing group with an extensive, well-managed infrastructure consisting of a distributed service architecture, a computer cluster of hundreds of nodes, and a secure network (http://genome.cgrb.oregonstate.edu/).

Aguilar, A., and J. C. Garza. 2006. A comparison of variability and population structure for major histocompatibility complex and microsatellite loci in California coastal steelhead (Oncorhynchus mykiss Walbaum). Mol Ecol 15:923-37. Araki, H. and M.S. Blouin. 2005. Unbiased estimation of relative reproductive success of different groups: evaluation and correction of bias caused by parentage assignment errors. Molecular Ecology, 13:4907-4110. Araki, H., W.R. Ardren, E. Olsen, B. Cooper and M.S. Blouin. 2007a. Reproductive success of captive-bred steelhead trout in the wild: evaluation of three hatchery programs in the Hood River. Conservation Biology, 21:181-190. Araki, H., B. Cooper and M.S. Blouin. 2007b. Genetic effects of captive breeding cause a rapid, cumulative fitness decline in the wild. Science, 318: 100-103. Araki, H., R.S. Waples, W.R. Ardren, B. Cooper and M.S. Blouin. 2007c. Effective population size of steelhead trout: influence of variance in reproductive success, hatchery programs, and genetic compensation between life-history forms. Molecular Ecology 16:953-966 Araki, H., R.S. Waples and M.S. Blouin. 2007d. A potential bias in the temporal method for estimating Ne in admixed populations under natural selection. Molecular Ecology 16: 2261–2271 Araki, H., B. Berejikian, M. Ford, and M.S. Blouin. 2008 Fitness of hatchery-reared salmonids in the wild. Evolutionary Applications 1:342-355. Araki, H., B. Cooper and M.S. Blouin. 2009. Carry-over effect of captive breeding reduces reproductive fitness of wild-born descendants in the wild. Biology Letters 5: 621-624. Arendt, J.D. 1997. Adaptive intrinsic growth rates: an integration across taxa. Quarterly Review of Biology 72:149-177. Arkush, K. D., A. R. Giese, H. L. Mendonca, A. M. McBride, G. D. Marty, and P. W. Hedrick. 2002 Resistance to three pathogens in the endangered winter-run chinook salmon (Oncorhynchus tshawytscha): effects of inbreeding and major histocompatibility complex genotypes. Can. J. Fish. Aquat. Sci. 59: 966-975 Baldwin, R., T.W. Miller, R.D. Brodeur,K.C. Jacobson. 2008. Expanding the foraging history of juvenile salmon: Combining stomach content and macroparasite community analyses for studying marine diets. Journal of Fish Biology 72:1268-1294. Berejikian, B., and M. J. Ford. 2004. Review of relative fitness of hatchery and natural salmon. National Oceanic & Atmospheric Administration Technical memo. Bernatchez, L., and C. Landry. 2003. MHC studies in nonmodel vertebrates: what have we learned about natural selection in 15 years? J Evol Biol 16:363-77. Bilby, R.E. and 9 co-authors. 2003. Review of salmon and steelhead supplementation. Report of the Independent Scientific Advisory Board to the Northwest Power Planning Council.(See page 30 of main report) http://www.nwcouncil.org/Library/isab/isab2003-3.htm Bird, A. 2007. Perceptions of epigenetics. Nature 447(7143): 396-398 Biro, P. A., M. V. Abrahams, J. R. Post, and E. A. Parkinson. 2004. Predators select against high growth rates and risk-taking behaviour in domestic trout populations. Proceedings: Biological Sciences 271:2233. Blouin, M. S. 2003. DNA-based methods for pedigree reconstruction and kinship analysis in natural populations. Trends in Ecology and Evolution 18:503-511. Blouin, M. S., M. Parsons, V. Lacaille, and S. Lotz. 1996. Use of microsatellite loci to classify individuals by relatedness. Mol Ecol 5:393-401. Blouin, M.S. V. Thuillier, B. Cooper, V. Amarasinghe, L. Cluzel, H. Araki and C. Grunau. 2010. No evidence for large differences in genomic methylation between wild and hatchery steelhead trout (Oncorhynchus mykiss). Canadian Journal of Fisheries and Aquatic Sciences. 67: 217-224. Brunelli, J. P., K. J. Wertzler, K. Sundin, and G. H. Thorgaard. 2008. Y-specific sequences and polymorphisms in rainbow trout and Chinook salmon. Genome 51:739-48. Christie, M., M.L. Marine and M.S. Blouin. Who are the missing parents? Grandparentage analysis identifies multiple sources of gene flow from hatchery fish into a wild population. Molecular Ecology, In review. Christie MR 2010. Parentage in natural populations: novel methods to detect parent-offspring pairs in large datasets. Molecular Ecology Resources 10(1), 115-128. Conover, D. O. 2004. Temperature-dependent sex determination in fishes. pp. 11 – 20. In: N. Valenzuela & V. Lance (editors) Temperature-dependent sex determination in vertebrates. Smithsonian Books, Washington, DC. Consuegra, S., and C. Garcia de Leaniz. 2008. MHC-mediated mate choice increases parasite resistance in salmon. Proc Biol Sci 275:1397-403. Craig JK, Foote CJ, Wood CC (1996). Evidence for temperature-dependent sex determination in sockeye salmon (Oncorhynchus nerka). Canadian Journal of Fisheries & Aquatic Sciences 53: 141–147. Derome, N., P. Duchesne, and L. Bernatchez. 2006. Parallelism in gene transcription among sympatric lake whitefish (Coregonus clupeaformis Mitchill) ecotypes. Molecular Ecology 15:1239. Dijkstra, J.M. et al.. 2007. A third broad lineage of major histocompatibility complex (MHC) class I in teleost fish; MHC class II linkage and processed genes. Immunogenetics, 59(4), 305-21. Evans, A. F., R. H. Wertheimer, M. L. Keefer, C. T. Boggs, C. A. Peery, and K. Collis. 2008. Transportation of Steelhead Kelts to Increase Iteroparity in the Columbia and Snake Rivers. North American Journal of Fisheries Management 28:1818-1827. Evans, M. L., B. D. Neff, and D. D. Heath. MHC genetic structure and divergence across populations of Chinook salmon (Oncorhynchus tshawytscha). Heredity 104:449-59. Filichkin, S. A., H. D. Priest, S. A. Givan, R. Shen, D. W. Bryant, S. E. Fox, W. K. Wong, and T. C. Mockler. Genome-wide mapping of alternative splicing in Arabidopsis thaliana. Genome Res 20:45-58. Fleming, I. A., and E. Peterson. 2001. The ability of released hatchery salmonids to breed and contribute to the natural productivity of wild populations. Nordic Journal of Freshwater Research 75:71-98. Ford, M. J. 2002. Selection in Captivity during Supportive Breeding May Reduce Fitness in the Wild. Conservation Biology 16:815. Forsberg, L. A., J. Dannewitz, E. Petersson, and M. Grahn. 2007. Influence of genetic dissimilarity in the reproductive success and mate choice of brown trout - females fishing for optimal MHC dissimilarity. J Evol Biol 20:1859-69. Fox, S., S. Filichkin, and T. C. Mockler. 2009. Applications of ultra-high-throughput sequencing. Methods Mol Biol 553:79-108. Fritts, A. L., J. L. Scott, and T. N. Pearsons. 2007. The effects of domestication on the relative vulnerability of hatchery and wild origin spring Chinook salmon (Onchorhynchus tshawytshca) to predation. Canadian Journal of fisheries and Aquatic Sciences 64:813-818. Frankham, R. 2008. Genetic adaptation to captivity in species conservation programs. Molecular Ecology 17:325-333. Gomez, D., P. Conejeros, S. H. Marshall, and S. Consuegra. MHC evolution in three salmonid species: a comparison between class II alpha and beta genes. Immunogenetics 62:531-42. Goodman, D. 2005. Selection equilibrium for hatchery and wild spawning fitness in integrated breeding programs. Canadian Journal of Fisheries & Aquatic Sciences 62:374. Grimholt, U., S. Larsen, R. Nordmo, P. Midtlyng, S. Kjoeglum, A. Storset, S. Saebo, and R. J. Stet. 2003. MHC polymorphism and disease resistance in Atlantic salmon (Salmo salar); facing pathogens with single expressed major histocompatibility class I and class II loci. Immunogenetics 55:210-9. Hansen, J. D., and S. La Patra. 2002. Induction of the rainbow trout MHC class I pathway during acute IHNV infection. Immunogenetics 54:654-61. Heath, D. D., J. W. Heath, C. A. Bryden, R. M. Johnson, and C. W. Fox. 2003. Rapid evolution of egg size in captive salmon. Science 299:1738-40. Hedrick, P. W. 2002. Pathogen resistance and genetic variation at MHC loci. Evolution 56:1902-8. Jacobson, K.C., D. Teel, D.M. VanDoornik and E. Casillas. 2008. Parasite associated mortality of juvenile Pacific salmon caused by the trematode Nanophyetus salmincola during early marine residence. Marine Ecology Progress Series 354:235-244. Jirtle, R.L., and Skinner, M.K. 2007. Environmental epigenomics and disease susceptibility. Nat. Rev. Genet. 8(4): 253-262 Johnson, N. A., R. L. Vallejo, J. T. Silverstein, T. J. Welch, G. D. Wiens, E. M. Hallerman, and Y. Palti. 2008. Suggestive association of major histocompatibility IB genetic markers with resistance to bacterial cold water disease in rainbow trout (Oncorhynchus mykiss). Mar Biotechnol (NY) 10:429-37. Keefer, M. L., R. H. Wertheimer, A. F. Evans, C. T. Boggs, and C. A. Peery. 2008. Iteroparity in Columbia River summer-run steelhead (Oncorhynchus mykiss): implications for conservation. Canadian Journal of Fisheries and Aquatic Sciences 65:2592. Kiryu, I. et al., 2005. New MHC class Ia domain lineages in rainbow trout ( Oncorhynchus mykiss ) which are shared with other fish species. Fish Shellfish Immunol. , 18, 243-254. Kostow, K. 2004. Differences in juvenile phenotypes and survival between hatchery stocks and a natural population provide evidence for modified selection due to captive breeding. Canadian Journal of Fisheries & Aquatic Sciences 61:577-589. Landry, C., D. Garant, P. Duchesne, and L. Bernatchez. 2001. 'Good genes as heterozygosity': the major histocompatibility complex and mate choice in Atlantic salmon (Salmo salar). Proc Biol Sci 268:1279-85. Launey, S., and D. Hedgecock. 2001. High genetic load in the Pacific oyster Crassostrea gigas. Genetics 159:255-65. Lohm, J., M. Grahn, A. Langefors, O. Andersen, A. Storset, and T. von Schantz. 2002. Experimental evidence for major histocompatibility complex-allele-specific resistance to a bacterial infection. Proc Biol Sci 269:2029-33. Lynch, M., J. Blanchard, D. Houle, T. Kibota, S. Schultz, L. Vassilieva, and J. Willis. 1999. Perspective: Spontaneous Deleterious Mutation. Evolution 53:645-663. Lynch, M., and M. O'Hely. 2001. Captive breeding and the genetic fitness of natural populations. Conservation Genetics 2:363-378. Mays, H. L., Jr., and G. E. Hill. 2004. Choosing mates: good genes versus genes that are a good fit. Trends Ecol Evol 19:554-9. McLean, J. E., P. Bentzen, and T. P. Quinn. 2005. Nonrandom, size- and timing-biased breeding in a hatchery population of steelhead trout. Conservation Biology 19:446-454. Miller, K. M., R. E. Withler, and T. D. Beacham. 1997. Molecular evolution at Mhc genes in two populations of chinook salmon Oncorhynchus tshawytscha. Mol Ecol 6:937-54. Montgomery, M. E., L. M. Woodworth, P. R. England, D. A. Briscoe, and R. Frankham. 2010. Widespread selective sweeps affecting microsatellites in Drosophila populations adapting to captivity: Implications for captive breeding programs. Biological Conservation 143:1842. Mortazavi, A., B. A. Williams, K. McCue, L. Schaeffer, and B. Wold. 2008. Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 5:621-8. Neff, B. D., S. R. Garner, J. W. Heath, and D. D. Heath. 2008. The MHC and non-random mating in a captive population of Chinook salmon. Heredity 101:175-85. Neff, B. D., and T. E. Pitcher. 2005. Genetic quality and sexual selection: an integrated framework for good genes and compatible genes. Mol Ecol 14:19-38. Ospina-Alvarez, N., and F. Piferrer. 2008. Temperature-dependent sex determination in fish revisited: prevalence, a single sex ratio response pattern, and possible effects of climate change. PLoS One 3:e2837. Pearsons, T. N., A. L. Fritts, and J. L. Scott. 2007. The effects of hatchery domestication on competitive dominance of juvenile spring Chinook salmon (Onchorhynchus tshawytscha). Canadian Journal of fisheries and Aquatic Sciences 64:803-812. Penn, D.J. 2002. The scent of genetic compatibility: sexual selection and the major histocompatibility complex. Ethology 108:1-21. Pitcher TE and BD Neff. 2007. Genetic quality and offspring performance in Chinook salmon: implications for supportive breeding. Conservation Genetics 8:607–616. Recover Implementation Science Team. 2009. A review of some applications of science to hatchery reform issues. April 9, 2009. http://www.nwfsc.noaa.gov/trt/puget_docs/hatchery_report_april92009.pdf Reik, W., and Walter, J. 2001. Genomic imprinting: parental influence on the genome. Nat. Rev. Genet. 2(1): 21-32. Reisenbichler, R., S. Rubin, L. Wetzel, and S. Phelps. 2004. Natural selection after release from a hatchery leads to domestication in steelhead, Oncorhynchus mykiss. Pp. 371-383 in B. R. Howell, E. Moksness and T. Svåsand, eds. Stock enhancement and sea ranching. Oxford, Malden, MA. Reisenbichler, R. R., and J. D. McIntyre. 1977. Genetic differences in growth and survival of juvenile hatchery and wild steelhead trout, Salmo gairdneri. Journal of the Fisheries Research Board of Canada 34:123-128. Reisenbichler, R. R., and S. Rubin. 1999. Genetic changes from artificial propagation of Pacific salmon affect the productivity and viability of supplemented populations. ICES Journal of Marine Science 56:459-466. Richards, E.J. 2006. Inherited epigenetic variation--revisiting soft inheritance. Nature Reviews Genetics 7(5): 395-401 Roberge, C., S. Einum, H. Guderley, and L. Bernatchez. 2006. Rapid parallel evolutionary changes of gene transcription profiles in farmed Atlantic salmon. Molecular Ecology 15:9. Seamons, T. R., P. Bentzen, and T. P. Quinn. 2004. The mating system of steelhead, Oncorhynchus mykiss, inferred by molecular analysis of parents and progeny. Environmental Biology of Fishes Shiina, T. et al., 2005. Interchromosomal duplication of major histocompatibility complex class I regions in rainbow trout (Oncorhynchus mykiss), a species with a presumably recent tetraploid ancestry. Immunogenetics, 56(12), 878-93. Steinberg, E. K., K. R. Lindner, J. Gallea, A. Maxwell, J. Meng, and F. W. Allendorf. 2002. Rates and Patterns of Microsatellite Mutations in Pink Salmon. Molecular Biology and Evolution 19:1198-1202. Spurgin, L. G., and D. S. Richardson. How pathogens drive genetic diversity: MHC, mechanisms and misunderstandings. Proc Biol Sci 277:979-88. Sundstrom, L. F., M. Lohmus, R. H. Devlin, and E. Brainerd. 2005. Selection On Increased Intrinsic Growth Rates In Coho Salmon, Oncorhynchus Kisutch. Evolution 59:1560. Tennessen, J. A., and M. S. Blouin. 2008. Balancing Selection at a Frog Antimicrobial Peptide Locus: Fluctuating Immune Effector Alleles? Mol Biol Evol 25:2669-2680. Tymchuk, W. E., L. F. Sundstrom, and R. H. Devlin. 2007. Growth And Survival Trade-Offs And Outbreeding Depression In Rainbow Trout (Oncorhynchus Mykiss). Evolution 61:1225. Waples, R. S., M. J. Ford, and D. Schmitt. 2007. Empirical results of salmon supplementation in the Northeast Pacific: A preliminary assessment. pp. 383-403 in T. M. Bert, ed. Ecological and Genetic Implications of Aquaculture Activities. Kluwer Academic Publishers. Wedekind, C., M. Walker, J. Portmann, B. Cenni, R. Muller, and T. Binz. 2004. MHC-linked susceptibility to a bacterial infection, but no MHC-linked cryptic female choice in whitefish. J Evol Biol 17:11-8. Whitehead, A., and D. L. Crawford. 2006. Variation within and among species in gene expression: raw material for evolution. Mol Ecol 15:1197-211. Ziegler, A., H. Kentenich, and B. Uchanska-Ziegler. 2005. Female choice and the MHC. Trends Immunol 26:496-502.

Review: RME / AP Category Review

Independent Scientific Review Panel Assessment

Assessment Number: 2003-054-00-ISRP-20101015
Project: 2003-054-00 - Evaluate the Relative Reproductive Success of Hatchery-Origin and Wild-Origin Steelhead Spawning Naturally in the Hood River
Review: RME / AP Category Review
Proposal Number: RMECAT-2003-054-00
Completed Date: 12/17/2010
Final Round ISRP Date: 12/17/2010
Final Round ISRP Rating: Meets Scientific Review Criteria
Final Round ISRP Comment:
This project was initiated in 2003 as one of a suite of projects to evaluate relative reproductive success of hatchery-origin steelhead compared to natural steelhead when spawning naturally, to address critical uncertainties identified in the 2000 FCRPS BiOp RPA 182.
First Round ISRP Date: 10/18/2010
First Round ISRP Rating: Meets Scientific Review Criteria
First Round ISRP Comment:

This project was initiated in 2003 as one of a suite of projects to evaluate relative reproductive success of hatchery-origin steelhead compared to natural steelhead when spawning naturally, to address critical uncertainties identified in the 2000 FCRPS BiOp RPA 182.

Documentation Links:
Proponent Response: