U.S. Dept. of Commerce / NOAA / OAR / PMEL / Publications


An Example of Fisheries Oceanography: Walleye Pollock in Alaskan Waters

Jim Schumacher

NOAA, Pacific Marine Environmental Laboratory, 7600 Sand Point Way NE, Seattle, WA 98115

Arthur W. Kendall, Jr.

NOAA, Alaska Fisheries Science Center, Seattle, Washington

U.S. National Report to International Union of Geodesy and Geophysics 1991–1994, Rev. Geophys., Suppl., 1153–1163 (1995)
Copyright ©1995 by the American Geophysical Union. Further electronic distribution is not allowed.

Introduction

A major area of research in fisheries oceanography examines relationships between recruitment dynamics of fish populations and the marine environment. A primary goal is to understand the natural causes of variability in year-class strength of commercially valuable species and apply this knowledge to management [Perry, 1994]. The paradigm that the majority of mortality occurs during transport of early life history stages from spawning to nursery grounds [Rothschild, 1986; Houde, 1987] provides an initial temporal focus for most research. The spatial domain includes the region occupied by early life history stages. Since global climate variability impacts regional ecosystem dynamics, however, the spatial domain often must be expanded. The relative importance and manifestation of biological factors (starvation and predation) that limit survival varies each year. Marked interannual and longer period variations in temperature (an influence on metabolic rates and behavior), transport of planktonic stages, and turbulence can exert an influence on both survival of early life history stages, and distribution of juveniles and adults. To understand how these environmental factors influence reproductive success of fish stocks also requires knowledge of the impact of these factors on predators and prey throughout the food web.

Our research, Fisheries-Oceanography Coordinated Investigations (FOCI, a NOAA program), examines recruitment dynamics of walleye pollock (Theragra chalcogramma) in Alaskan ecosystems [Reed et al., 1988; Schumacher and Kendall, 1991; Reed et al., 1994]. Fisheries oceanography is a very broad field of research with many programs and various approaches; FOCI represents a subset of this research. FOCI began in 1984 in Shelikof Strait, Alaska; in summer 1991 another element (a component of NOAA's Coastal Ocean Program) began in the Bering Sea. Pollock constitute the world's largest single species fishery with annual catches from Alaskan waters exceeding 1.2 million metric tonnes [Westpested, 1993]. In both regions, large variations in recruitment (> twenty-fold) dictate management decisions on harvest quotas. Understanding these variations requires knowledge of what conditions result in survival: each individual must successfully navigate through a sequence of hurdles that likely are coupled in nonlinear ways, vary each year, and no single condition accounts for a large fraction of the observed interannual signal in recruitment.

The inherent complexity of the ecosystem requires research specialization, which then must be integrated to provide a useful product. A better understanding of the population dynamics process requires more interdisciplinary research among fisheries scientists and oceanographers [Beamish et al., 1989]. Unfortunately, such varied backgrounds traditionally result in disparate rather than integrated research [Wooster, 1986]. The tendency to do research without being responsible for implementation of results also detracts from achieving holistic goals. It has long been known that humans impact their environment. "Man did not weave the web of life; he is merely a strand of it. Whatever he does to the web, he does to himself" (Chief Seattle, 1854). Integrating research conducted by people from disparate academic backgrounds and maintaining a goal of responsible implementation of results can ensure that fishing mortality does not detrimentally interact with natural fluctuations in recruitment. In this way we acknowledge the gifts nature provides and leave a legacy for future generations. This is not only our moral responsibility, it is mandated by Federal law (Magnuson Fisheries Management and Conservation Act, 1976).

In this overview, we present some of the major developments and results from FOCI. First, we present background information for both FOCI programs (Shelikof Strait and Bering Sea), including research strategy and regional circulation features. We then present results that couple the biology and physics and are important features in both programs, as are the methods and techniques that follow. Finally, we present the application of research results from Shelikof Strait FOCI to management. The goal of FOCI, to understand natural fluctuations in year-class strength of pollock and to provide information to reduce uncertainty in status-of-pollock-stock models, is applied science. The initial examination of the physical and biological environment and the time/space distribution of various life history stages of pollock is viewed as fundamental science. Development and implementation of models to integrate biophysical observations, and technologies to measure conditions, as well as developing methods to apply new findings to management occur as FOCI matures.

Shelikof Strait FOCI

Research Strategy

Prior to FOCI, only very limited knowledge existed of either the physical environment or the life history of pollock in the western Gulf of Alaska. What was known suggested the following hypothesis: optimum survival and subsequent recruitment result when larvae are transported to nursery grounds in coastal regions along the Alaska Peninsula, rather than into the Gulf of Alaska. Further, biophysical processes occurring during transport have significant impact on survival. Initially studies focussed on the life history of pollock and the nature of the regional circulation. As these aspects became known, field operations switched to maintenance time series of selected biological and physical characteristics, studies of biophysical processes, and development of methods to examine preferential survival. Simultaneously, we began implementation of coupled biophysical and correlative models, and development of methods to transfer FOCI results to the assessment and prediction models used to provide biomass scenarios to management.

Pollock Life History

Separation of pollock larvae from those of similar species in field samples was first accomplished in 1981 [Dunn and Matarese, 1987]. At the same time, the strong year-classes in the mid-1970's resulted in a large fishery [Megrey, 1990], and hydroacoustic surveys were conducted to estimate the adult population for input to assessment and prediction models. The surveys showed that large concentrations of pre-spawning walleye pollock migrate from the southwest end of Shelikof Strait (taken to include the sea valley that extends from the Barren Islands southwestward to the slope off Chirikof Island, Figure 1) primarily to the area near Cape Kekurnoi. Spawning takes place mainly in the deep sea valley in early April [Kim and Nunnallee, 1991], at depths between 150-250 m. Each female produces about ½ million free floating planktonic eggs in a series of about 10 batches over a period of a few weeks. The eggs reside mainly below 150 m and hatch in about 2 weeks. The short localized spawning pattern creates a large "patch" of eggs observed in plankton surveys [Kendall and Picquelle, 1990]. Mortality rates of the eggs decrease through the spawning season from 0.4/day to 0.1/day [Kim and Gunderson, 1989].

Figure 1. Upper Panel: The FOCI study area in the western Gulf of Alaska with a schematic showing the location of pollock life history stages. The outlined arrows represent potential transport of larvae off the shelf. Lower Panel: Circulation in the study area as derived from satellite-tracked buoys (51) drogued at 40 m and deployed between 1986 and 1994 (after Schumacher and Kendall [1991]).

The larvae are about 3-4 mm Standard Length (SL) at hatching and are relatively undeveloped, without functioning mouths or eyes. They quickly rise from their deep hatching depth to the upper 50 m of the water column, where they drift in the prevailing currents for the next several weeks (late April through May). They generally remain in identifiable patches, growing about 0.2 mm/day with a daily mortality rate of ~8.7% during this time. Their diet consists mainly of early life stages of copepods [Canino et al., 1991], and the size range of prey increases as the larvae grow. Larger larvae undergo diel migrations (deepest during the day) between 15-50 m. Larvae appear to stay just below the turbulent wind mixed upper layer of the water column [Kendall et al., 1994]. They are visual feeders, eating mostly during the day [Kendall et al., 1987].

By late May considerable year-to-year variation exists in abundance and location of larvae [Kendall and Picquelle, 1990] as well as the zooplankton that produce their prey [Incze and Ainaire, 1994]. Two primary tracks characterize westward larval drift: along the Alaska Peninsula in the relatively slow moving flow (<10 cm s-1) over the shoreward edge of the sea valley, and offshore in the rapidly (>25 cm s-1) moving Alaska Coastal Current [Kim and Kendall, 1989]. By midsummer the larvae have transformed into juveniles, which by late summer are schooled and become concentrated in nearshore areas along the Alaska Peninsula [Hinckley et al., 1991]. After this period, we know little about their life until they enter the fishery and become sexually mature at age ~3 years. The adults live for ~11 years.

We have established the life history pattern between spawning migration and young of the year (Figure 1). Using results from egg and larval surveys, together with estimates of the spawning adult population from hydroacoustic observations and indices of juvenile abundance [Bailey and Spring, 1992], the stages when year-class strength is established have been determined. A low abundance of larvae results in weak recruitment. A large abundance of larvae, however, does not imply strong recruitment. This suggests that processes during the juvenile stage can also be critical [e.g., Bailey et al., 1994b]. Recently, we have focussed more research on understanding young of the year juveniles since interannual variability in their survival also affects year class strength [Bailey and Spring, 1992].

Regional Circulation and Mesoscale Features

Much of the variability in the physical environment in the Gulf of Alaska results from large scale atmospheric phenomena. Global patterns in the upper level atmospheric pressure generate climatic conditions that include an annual cycle in the number of low pressure centers traversing the region [Niebauer, 1988]. The consistent passage of storms along the Aleutian Island chain (the Aleutian Low) dominates wintertime atmospheric circulation. Interaction of frequent storms with the mountainous coastline results in a high precipitation rate (>200 cm yr-1) along the coastal region. The discharge rate of freshwater reflects seasonal variations in air temperature, precipitation, runoff and storage from the previous winter [Royer, 1982].

Along the Alaska Peninsula, along with a deep (>250 m) sea valley, a high, nearly continuous mountain chain exists (Figure 1). The mountains perturb regional winds so that in Shelikof Strait proper down-gradient winds are common [Schumacher et al., 1989], and the winds over coastal waters west of Kodiak Island are altered for ~60 km offshore [Macklin et al., 1993].

The dominant circulation feature is the Alaska Coastal Current (ACC), a distinct flow that only 20 years ago was unknown. FOCI research has elucidated many of the characteristics of the ACC, which extends for >1500 km along the coast of Alaska [Reed and Schumacher, 1981]. This is one of the most vigorous coastal currents in the world with speeds typically between 25 and 100 cm s-1 [Stabeno et al., 1995]. Volume transport results from the addition of freshwater along the entire coastline and is perturbed by the alongshore wind through both confinement of the freshwater and alteration of coastal sea level [Schumacher and Reed, 1980; Royer, 1981; Reed and Schumacher, 1981]. The observed mean transport in Shelikof Strait is ~0.80 × 106 m3 s-1; wind forced pulses exceed 3.0 × 106 m3 s-1 [Schumacher et al., 1989; Stabeno et al., 1995]. Wind-driven fluctuations within the strait proper are greater than those over coastal waters east of Kodiak Island due to the topographic effects on the winds [Stabeno et al., 1995]. Differential Ekman pumping may amplify this mechanism within the strait proper [Reed and Schumacher, 1989a, b]. Estimates of net volume transport computed from water property observations collected between 1985 and 1992 have a mean of 0.66 × 106 m3 s-1 [Reed and Bograd, 1995].

In Shelikof Strait, horizontal density gradients and vertical shear in the mean flow create the baroclinic instability, evident in satellite images [Vastano et al., 1992; Schumacher et al., 1991] and in analysis of current records [Mysak et al., 1981], which dominates flow patterns and generates eddies [Schumacher et al., 1993]. The ACC does not span the sea valley, and estimates of coherence become insignificant for separations >10 km [Reed and Schumacher, 1989b; Bograd et al., 1994]. Estuarine-like flow also exists, with warmer more saline water from the continental slope entering on the southeastern side of the valley [Reed et al., 1987]. The ACC bifurcates east of Sutwik Island; one branch continues along the Alaska Peninsula and the other flows seaward through the sea valley [Schumacher et al., 1989]. Since 1986, 51 satellite tracked buoys (drogued at 40 m) were deployed in the study area during spring near Cape Kekurnoi. To date, 25% of the buoys continued along the Peninsula. The remainder moved seaward past the Semidi Islands, most, however, traveled shoreward (between ~157° and 158°W) and joined the flow along the Peninsula. Only 25% of the buoys left the shelf permanently and became incorporated in the Alaskan Stream [Stabeno and Reed, 1991].

FOCI research has found and elucidated the dominant circulation and mesoscale features: the ACC, eddies generated by baroclinic instability, and an estuarine-like flow of slope waters into the sea valley. These features must be simulated in any numerical model of the region. Further, results show that the timing and location of hatching determines whether larvae enter an eddy, or are transported with either the slow-moving coastal flow or the rapid ACC. Modeling studies [Stabeno et al., 1995] suggest that the location of late larvae varies greatly year to year depending on advection. The phasing between biological and physical processes determines transport of larvae and presumably their eventual recruitment.

Bering Sea FOCI

Research Strategy

Recommendations from an International Symposium on Pollock [Aron and Balsiger, 1989] provide the research objectives for FOCI: determine stock structure in the Bering Sea and its relationship to physical features, and understand recruitment processes in the eastern Bering Sea. Both of these have direct implication to management of the vast resources that exist in U.S., Russian and international waters. To attain the first objective, field and modeling studies have investigated circulation throughout the deep basin. Another component seeks to establish genetic "finger-prints" to evaluate stock structure. In addressing the second objective, we are investigating differences between survival of eggs and larvae over the deep waters to that over the adjacent shelf. A newly established component is comparing habitats of juvenile animals around the Pribilof Islands.

Pollock Life History

Many of the characteristics of walleye pollock early life history are common to all populations of the species. In the Bering Sea, however, both the population structure and early life history pattern are much more complex than in the Gulf of Alaska. Genetic characteristics [Mulligan et al., 1992] and length-at-age and fecundity relationships [Hinckley, 1987] suggest several spawning stocks exist. The importance of pollock in the ecosystem [e.g., Springer, 1992], as well as the relationships and interchange among stocks are largely unknown. Spawning begins earlier in the year in some parts of the Bering Sea than it does in Shelikof Strait and apparently different groups of fish spawn at different times and places. We began our efforts focusing on the population that spawns in February over the southeastern slope, and supported a substantial fishery in the late 1980's. Here we found indications that some of the eggs and larvae were much deeper in the water column (400 m) than we had found in Shelikof Strait. Also, feeding conditions did not seem to be adequate for optimal growth in this area. We found larvae associated with eddies in this area as we had in Shelikof Strait. Our attention is now focused on the spawning (April-June) that occurs over the continental shelf of the southeastern Bering Sea.

Basin Circulation and Mesoscale Features

Prior to FOCI research many schematics existed of circulation in the Bering Sea, and wind stress was considered to provide the primary forcing [Hughes et al., 1974]. Results from FOCI have refined our knowledge of circulation (Figure 2) and meteorological forcing over the basin from both observations [Stabeno and Reed, 1994] and model studies [Overland et al., 1994]. A cyclonic gyre dominates circulation over the basin, with a western boundary current (Kamchatka Current) along the Asian side of the basin [Reed et al., 1993]. This gyre is mainly an extension of the Alaskan Stream, and the majority of volume transport enters through Near Strait (~10 × 106 m3 s-1) and exits via the Kamchatka Current [Stabeno and Reed, 1994]. When instabilities in the Alaskan Stream inhibit flow into the Bering Sea through Near Strait [Stabeno and Reed, 1992], transport in the Kamchatka Current can be reduced by ~50%. Such conditions existed from 1990 to 1991; the return to normal flow conditions occurred in late 1991 [Reed and Stabeno, 1993]. A climatology of the wind forcing shows that eastward- and northward-propagating storm systems dominate the surface stress at short periods (<1 month), which serves principally to mix the upper ocean [Bond et al., 1994]. At longer periods (>1 month), the estimated wind-driven transport accounts for roughly one-half of the observed transport within the Kamchatka Current. The interannual variations in the transports are ~25% of the mean.

Figure 2. Upper Panel: A schematic of the general circulation over the basin of the Bering Sea as derived from ongoing FOCI research (after Stabeno and Reed [1994]). Lower Panel: A schematic of circulation over the eastern shelf based on previous results [Schumacher and Kinder, 1983], together with more recent satellite-tracked buoy [Stabeno and Reed, 1994] and moored current observations [Schumacher and Reed, 1992]. ACC is the Alaska Coastal Current, which enters through Unimak Pass, and W represents regions with weak or statistically insignificant mean flow.

The flux (~3.0 × 106 m3 s-1) of Alaskan Stream waters through the eastern passes (Amchitka and Amukta Passes) has a profound impact on regional water properties and circulation [Schumacher and Stabeno, 1994; Reed et al., 1994; Reed and Stabeno, 1994]. These waters then flow northwestward along the slope [Schumacher and Reed, 1992], carrying a subsurface temperature maximum that can be traced hundreds of kilometers. The southeastern basin waters are also rich in eddies, some of which are formed by flow through Amukta Pass [Schumacher and Stabeno, 1994].

Coupled Biophysical Processes

We use the term coupled biophysical processes to mean physical processes that influence biological conditions affecting various life history stages of pollock.

Eddies

In both Shelikof Strait and the Bering Sea, eddies play a role in coupled biophysical processes. These features frequently occur over the sea valley west of Kodiak Island, as revealed in infrared, synthetic aperture radar, and color scanner satellite imagery [Reed et al., 1988; Schumacher et al., 1991; Vastano et al., 1992; Liu et al., 1994], buoy trajectories [Incze et al., 1990], water property and larval distributions [Schumacher et al., 1993], and moored current records [Bograd et al., 1994]. The location of eddy formation coincides with the spawning region. Formation of three or four eddies per month during spring [Bograd et al., 1994] assures that some eggs hatch into an eddy. As a result of limited dispersion, high abundances of larvae often exist in such features. Further, since some eddies tend to remain nearly stationary for periods of weeks, they retain larvae in the sea valley [Reed et al., 1989; Vastano et al., 1992; Schumacher et al., 1993]. Some eddies have unique chemical properties [Incze et al., 1989], which may aid prey production and survival of larvae in the eddy [Schumacher and Kendall, 1991]. Nutritional condition of first-feeding larvae can be reflected by the ratio of ribonucleic acid to deoxyribonucleic acid (RNA/DNA). Concentrations of nauplii, larval gut contents, and RNA/DNA were higher for larvae in an eddy than for those in adjacent waters [Canino et al., 1991; Bograd et al., 1994].

Results from observations during May 1990 showed a connection between an eddy and larvae [Schumacher et al., 1993]. Contours of larval abundance coincide with those of salinity, and lie in close proximity to buoy trajectories. Physical data showed little or no exchange of water between the eddy and adjacent waters, permitting estimates of mortality that reflect only predation and/or starvation. The observations yield a daily mortality (4.4%) low compared to other estimates based on observations (6.3%, Reed et al. [1989]; 5-11%, Yoklavich and Bailey [1990]), or from dispersion model simulation (5.9-8.0%, Kim and Bang [1990]).

In the eastern Bering Sea, the presence of relatively small eddies (diameter <50 km) has recently been documented [Schumacher and Stabeno, 1994]. These are formed in regions of high current shear in open waters, or by interaction of inflowing Alaskan Stream water with topography of passes in the eastern Aleutian Island chain. Eddy formation during periods of inflow through this pass occur in a numerical model of the region (M. Spillane, pers. comm.). Since 1986, 58 satellite-tracked buoys have been deployed in the Bering Sea to support studies of pollock and their environment. In 4 of these years, five regions of high rough counts of pollock larvae were found and buoys deployed in them. In all but one case, the trajectories of the buoys defined eddies. Likewise, the buoys (33) that were not deployed in a patch did not indicate eddies. This association of pollock larvae and eddies may have significant impact on larval survival.

Turbulence

General results from modeling studies suggest that wind mixing of the upper water column can both be beneficial or detrimental to larval survival, depending on the intensity of the turbulence [Davis et al., 1991]. Many processes exist that connect pollock larval survival to mixing [Bailey and Macklin, 1994]. These authors determined a time series of abundance of larvae hatched on a given day that survived through the early feeding stage, and established a mixing index using the cube of the wind speed. Comparing these series revealed two patterns: strong wind events during the first-feeding period coincided with lower than expected survival, and periods of higher than expected larval survival were associated with calm periods of wind often bracketed by strong mixing. During a spring with moderate winds and a shallow mixed layer, concentrations of food, growth at age and mortality rates were more conducive to larval survival than during a spring when strong winds were accompanied by a deep mixed layer [Bailey et al., 1994b].

Bailey and Macklin suggest how larval survival may be related to an integration of wind-mixing, stratification within an eddy, and larval behavior. The eddy observed in 1989 had enhanced prey and feeding conditions and a low-salinity core relative to surrounding waters. Pollock larvae in the laboratory avoid turbulence [Olla and Davis, 1990] by moving deeper in the water column. Reduced light intensity with increasing depth has detrimental effects on the ability of larvae to search for and capture prey [Heath, 1989]. The vertical stratification of the eddy required more wind-induced turbulence than adjacent waters to mix to comparable depths. Thus, under similar winds, larvae within the eddy could remain higher in the water column in better feeding conditions than larvae outside the eddy. Hence, first-feeding larvae are more likely to survive in the eddy than in the surrounding waters giving similar prey fields.

Methods and Techniques

Models

We have used models in many ways, including to conceptualize program strategies, to expand the time-space domain of field observations, to guide program direction through hypothesis generation, to elucidate potential linkages among sets of biological and physical factors, and to provide information into the fisheries management stream.

Shelikof Strait Physical/Biological Model

Ongoing modeling studies examine potential impact of interannual changes in circulation on survival of larvae in the western Gulf of Alaska. Physical factors that pose challenges include complex bathymetry with many islands, mesoscale (~20 km) meanders and eddies, strong vertical shear (estuarine-like flow), and strong forcing by winds and freshwater runoff. The circulation model we use is based on the Semispectral Primitive Equation Model (SPEM) of Haidvogel et al. [1991], modified for this region [Stabeno et al., 1995]. A total of ~250,000 gridpoints span the model domain, and a typical simulation of spring and summer months entails ~105 time steps. The model code is presently designed to take advantage of vector-processing computing architectures.

Thus far SPEM has reproduced the observed general spatial features of circulation [Stabeno et al., 1995]. A comparison between model output and measured currents yielded reasonable agreement. The model also generates eddies with similar spatial scales to those observed. Results from SPEM show that during 1978 (the strongest year class) larvae were more likely transported into coastal waters along the Alaska Peninsula, while in 1990 (a below average year class) they remained in the sea valley where currents then result in transport offshore [Stabeno et al., 1995]. This latter scenario implies loss of recruits. These results support the original transport hypothesis.

SPEM is being coupled to a spatially explicit, individual-based, probabilities model (IBM) of egg and larval development. The IBM has distinct advantages over more traditional approaches that consider only the "mean" individual [Huston et al., 1988]; since it follows the unique life history of each fish, the IBM approach yields specific information about survivors. The model employs a spatial tracking algorithm for each individual, that includes vertical migration according to life stage. Horizontal transport, growth, and behavior are governed by velocity, salinity and temperature fields generated by SPEM. Low-pass filtered velocity and scalar fields from SPEM are stored once per model day, then used as input for multiple runs of the biological model. The model-generated spatial distributions qualitatively compare favorably with observed distributions of larvae and juveniles. Interannual differences in wind and freshwater runoff lead to differences in the modeled spatial paths of individuals, and in the distributions of population attributes (e.g., growth).

Coupled 1-dimensional Physical Biological Model

The objective of this component is to adapt and extend a coupled physical biological 1-D model to investigate production dynamics of the pelagic ecosystem as it pertains to survival of larval pollock. The approach uses field observations to determine rates and appropriate species composition for several of the distinct physical [Coachman, 1986] and biological [Smith and Vidal, 1986] domains in the eastern Bering Sea. The present model includes stage-structured dynamics of copepod populations (Calanus and Neocalanus) and larval pollock feeding and growth (S. Bollens, pers. comm.). The temporal behavior of the mixed layer comes from observations (see below). Model results include that the species composition of zooplankton has a strong influence on growth of larval pollock; the presence of protozoan prey becomes important when young copepods are scarce (i.e., early spring over the slope). Results from the model also suggest that variability of the mixed-layer depth has significant impact on larval growth by affecting lower trophic level production. These results have led to the addition of a research component that addresses the potential importance of protozoans as prey for larval pollock.

Correlative Models

Many studies attempt to examine the causes of recruitment variation in fishes and relate these to biotic and abiotic factors. Hollowed [1992] lists 47 such studies along the coast of the northeast Pacific. There are two schools of thought regarding the utility of the correlative approach. It can be considered futile because of biases, measurement error, and the near certainty of spurious correlations [Walters and Collie, 1988]. Others claim that these studies provide information on patterns that lead to testable hypothesis [Kope and Botsford, 1990]. Recent studies [Tyler, 1992; Hollowed, 1992] advocate the use of correlative studies with the constraint that the analysis be based on a sound conceptual framework and judicious use of statistical methods. This approach was applied to a recruitment time series for adult pollock for the period of 1962-1989 and egg and larval series for 1981-1989 [Megrey et al., 1995]. The physical time series included estimates of precipitation, sea-level atmospheric pressure, wind-driven mixing, transport and water properties. Statistical techniques suggest that recruitment, as well as indices of age-0 and age-1 abundance, were related to precipitation, an index of sea level pressure, and wind mixing. Given the time period and phasing of the physical factors, the results imply that stratification can influence behavior of and predation on juveniles, and that increased baroclinicity influences larval survival through its impact on eddy generation and wind mixing. The former implication supports previous studies of juvenile behavior [Olla and Davis, 1995], while the latter two have been inferred from analysis of biophysical observations.

Biophysical Observations from a Moored Platform

In order to examine biophysical conditions from pre-spring bloom conditions through the summer, we have deployed a mooring over the outer slope (~2200 m). The platform itself represents a "hardened" version of moorings along the equator [McPhaden et al., 1991]. Included in the suite of measurements are winds, insolation, air temperature, humidity, salinity/temperature/pressure (at ten depths), currents (both Acoustic Doppler Current Profiler and acoustic current meters at depth), acoustic backscatter, and chlorophyll absorption [Moore et al., 1992]. Some of the observations are sent real time via a satellite system. When a significant alteration of the diel migration occurred, the availability of real-time data permitted direction of field sampling to provide in situ measurements. Results to date have elucidated characteristics of eddies, including both their density and velocity structure. Further, the observations provide time series of mixed layer depth to a coupled 1-dimensional, physical-biological model.

Real-Time Detection of Eddies

Results from both Shelikof Strait and the slope waters of the eastern Bering Sea indicate that the highest abundances of pollock larvae often reside in eddies. To examine the nature of biophysical processes extant in these features and determine their influence upon survival requires in situ observations. Finding a reliable method to locate an eddy for field studies provides a challenge. Although infrared imagery has proved useful, cloud cover and generally weak sea surface temperature gradients limit this approach. High resolution Synthetic Aperture Radar (SAR) eliminates both of these constraints. Mesoscale features are imaged by SAR through several possible mechanisms that are not well understood, including modulation of the short surface waves by current shear, alteration of the stability of the surface wind across a relatively sharp sea surface temperature gradient, surface film damping of short surface waves, shifts in the Doppler frequency due to variability in the surface currents, and current-induced wave refraction. This latter mechanism has been examined for features in Shelikof Strait [Liu et al., 1994].

During April and early May 1992, three eddies (20-25 km diameter) were apparent in SAR images of Shelikof Strait. In mid-May a larval survey was conducted in the same region. A satellite-tracked buoy deployed in a region of high larval abundance made a circular trajectory around a mesoscale feature that likely was one of the eddies observed in the SAR imagery [Liu et al., 1994]. During this and a subsequent cruise, anomalous patterns of backscattering appeared on a 38-kHz acoustic system. A strong scattering layer at the surface and in midwater, with the column in between nearly void of sound scattering organisms, characterized the signal. This signal appeared in several sections of data where SAR had indicated the presence of eddy-like features. Analysis of concomitant water property and shipboard acoustic Doppler Current Profiler (150 kHz) observations confirmed the existence of these features. The density of larval pollock in these features was estimated to be an order of magnitude greater than in surrounding waters. Acoustic backscatter signals can sometimes be used to identify and characterize mesoscale biophysical features in the ocean [Aoki and Inagaki, 1992], thereby permitting real-time studies of these features.

Immunoassay, Otolith, and RNA/DNA Techniques

It has long been thought that processes involving nutrition and predation of larval fishes play significant roles in their survival and ultimately in the strength of year classes. Until recently, however, lack of suitable technology has hampered efforts to study these processes. The recent discovery of a record of daily growth of larvae in their otoliths (ear bones) has been applied to estimate growth and survival rates of a large number of species [Campana and Neilsen, 1985; Jones, 1986]. Immunoassay techniques, in which antibodies to particular fish species are developed and used to detect the presence of macerated remains of eggs or larvae of that species in predator guts, have also found increasing use [Theilacker et al., 1993]. Determining the nutritional state of larvae through histological assessment, and determining RNA-DNA ratios of whole larvae, or in cells of particular organs, has proven very valuable [Theilacker, 1978; Buckley, 1984; Clemmensen, 1988; Theilacker and Shen, 1993; Canino, 1994]. All of these techniques have been refined and applied in FOCI studies.

Laboratory rearing studies with larval pollock confirmed that increments are deposited daily in their otoliths [Bailey and Stehr, 1988]. Growth rates and hatch dates, based on length and age determined by otolith increments, of field-collected pollock larvae from various years and areas, have been compared [Yoklavich and Bailey, 1990]. Growth rates were found not to vary interannually, but the hatching period did. By estimating decreases in the abundance of cohorts within year classes with time during the larval period, mortality rates have been calculated [Yoklavich and Bailey, 1990]. These techniques were used to investigate differences in larval survival during the season. It was found that larvae that were at a first feeding stage during calm weather, had higher survival rates than larvae that reached first feeding during storms [Bailey and Macklin, 1994]. Growth rates of young-of-the-year juveniles have been determined, although beyond the larval period otolith growth is more complex and grinding is required to discern all of the increments [Brown and Bailey, 1992; Bailey et al., 1994a]. Besides using the record of daily growth in larval otoliths, studies have analyzed the elements deposited in the otoliths as a record of the environment experienced by the larvae at various times during their development. This technique has been investigated as a means of discriminating among juvenile pollock of various geographic/genetic origins [Severin et al., 1995].

To investigate predation on fish eggs and larvae by invertebrates that macerate their prey, antibodies against yolk proteins were developed, and potential predator gut contents have been assayed for the presence of these proteins [Bailey et al., 1993; Brodeur and Merati, 1993]. Decapod shrimp, euphausids, and gammarid amphipods were all found to consume significant numbers of pollock eggs and yolk-sac larvae. We are now developing immunoassays for later stage larvae to investigate the role of predators in their mortality (Brodeur, R.D. and N. Merati, pers. comm.).

DNA of individual cells is fairly constant, so the DNA content of whole animals increases proportional to increases in cell number (growth). However, RNA content of cells is variable, and reflects active protein synthesis. RNA/DNA ratios have been found to be accurate indicators of recent feeding of larvae: higher ratios indicate better feeding condition [Canino, 1994]. Condition of pollock larvae, measured by RNA/DNA in whole individual larvae, has been found to vary with location, time of year, and interannually [Canino et al., 1991; Bailey et al., 1994b]. Variations in larval condition were concordant with variations in prey abundances. Thus it appears that pollock larvae in Shelikof Strait may experience feeding conditions that may limit growth. A prolongation of the larval stage due to reduced growth rate likely increases mortality. Even more precise indicators of larval nutritional condition and recent growth history have been developed using flow cytometry to measure RNA and DNA contents of brain cells of individual larvae [Theilacker and Shen, 1993]. These methods can be applied to assay larvae at sea in near real time.

Information Transfer to Fisheries Management

NOAA's National Marine Fisheries Service (NMFS) advises the North Pacific Fisheries Management Council on the status of pollock stocks in the Gulf of Alaska and Bering Sea. The process includes a stock projection model, initialized with results from a model of stock assessment. The latter model employs commercial catch statistics and operational survey information to produce estimates of stock abundance. Given these results, the stock projection model forecasts future abundance as a function of different harvest and recruitment scenarios. The stock assessment model requires information independent of the fishery to calibrate estimates of absolute population. Prior to FOCI guidance, these were limited to the hydroacoustic and bottom-trawl surveys. For the last 2 years, estimates of spawning biomass for the Gulf of Alaska derived from the Annual Egg Production Method [Picquelle and Megrey, 1993] have provided a third source of information.

The effects of management decisions on the Shelikof Strait pollock population are examined using alternative harvest and recruitment scenarios in the stock projection model. Beginning in 1992, FOCI has analyzed biological and physical time series to estimate recruitment qualitatively. This prediction significantly simplifies the stock projection analysis by limiting the number of viable recruitment scenarios. The forecast of year-class strength uses hydroacoustic survey results of spawning aggregations, ichthyoplankton surveys of eggs and larvae, estimates of spawning biomass/recruitment from the annual stock assessment [Hollowed et al., 1993], and a suite of biological and physical factors. Previous modeling of pollock recruitment [Megrey et al., 1995] did not address the autocorrelated nature of the recruitment data. In 1993, the recruitment modeling was augmented with the development of a transfer function model. This model uses the same physical environmental data as before, but accounts for the autocorrelation of recruitment. In addition, it directly predicts recruitment. In 1993 it was used to generate recruitment scenario candidates for the stock projection model. FOCI's prediction made in 1992 was for weak 1989-90 year classes, a weak to average 1991 year class, and a strong 1992 year class. Recent observations of recruitment to the 1989 and 1990 year classes are consistent with this prediction.

Acknowledgments. We thank the numerous scientists who planned, conducted and published their research, and those who have provided information not yet published. In particular, Phyllis Stabeno, Ron Reed, and Jeff Napp for their continuous support and critical comments on this manuscript, and Allen Macklin, FOCI Program Coordinator. We also thank all the technical staff who assisted with the data acquisition, preparation, and analysis. Continued superlative support has been provided by the NOAA Ship Miller Freeman, and on specific cruises by the NOAA Ships Discoverer and Surveyor. A portion of this research and analysis was funded by Minerals Management Service, Interagency Agreement #14-35-0001-14165. The majority of research was funded by Fisheries Oceanography Coordinated Investigations and NOAA's Coastal Ocean Program. Wopila Tunkashila, Maki Unci and Grandmother Ocean. This is FOCI contribution #212 and Pacific Marine Environmental Laboratory contribution #1535.

References

Aoki, I., and T. Inagaki, Acoustic observations of fish schools and scattering layers in a Kuroshio warm-core ring and its environs, Fish. Oceanogr., 1, 137-142, 1992.

Aron, W., and J. Balsiger, Proceedings of the International Scientific Symposium on Bering Sea Fisheries, NOAA Tech. Memo. NMFS F/NBC-163, 424 pp., 1989.

Bailey, K. M., R. D. Brodeur, N. Merati, and M. M. Yoklavich, Predation on walleye pollock (Theragra chalcogramma) eggs and yolk-sac larvae by pelagic crustacean invertebrates in the western Gulf of Alaska, Fish. Oceanogr., 2, 30-39, 1993.

Bailey, K. M., A. L. Brown, A. Nishimura, and M. T. Reilly, Tools for 3D-imaging of walleye pollock Theragra chalcogramma otoliths: reconstruction from serial sections and fluorescent laser cytometry, J. Fish. Biol., in press a.

Bailey, K. M., M. F. Canino, J. M. Napp, S. M. Spring, and A. L. Brown, Two contrasting years of prey levels, feeding condition and mortality of larval walleye pollock (Theragra chalcogramma) in the western Gulf of Alaska, Mar. Ecol. Prog. Ser., in press b.

Bailey, K. M., and S. A. Macklin, Analysis of patterns in larval walleye pollock Theragra chalcogramma survival and wind mixing events in Shelikof Strait, Gulf of Alaska, Mar. Ecol. Prog. Ser., 113, 1-12, 1994.

Bailey, K. M., and S. Spring, Comparisons of larval, age-0 juvenile and age-3 recruit abundance indices of walleye pollock Theragra chalcogramma in the western Gulf of Alaska, ICES J. Mar. Sci., 49, 297-304, 1992.

Bailey, K. M., and C. L. Stehr, The effects of feeding periodicity and ration on the rate of increment formation in otoliths of larval walleye pollock Theragra chalcogramma (Pallas), J. Exp. Mar. Biol. Ecol., 122, 147-161, 1988.

Beamish, R. J., G. A. McFarlane, and W. S. Wooster, The need for interdisciplinary research in fisheries and ocean sciences, in Effects of Ocean Variability on Recruitment and an Evaluation of Parameters Used in Stock Assessment Models, edited by R. J. Beamish and G. A. McFarlane, Can. J. Fish. Aquat. Sci., Spec. Pub.,108, pp. 1-3, 1989.

Bograd, S. J., P. J. Stabeno, and J. D. Schumacher, A census of mesoscale eddies in Shelikof Strait, Alaska during 1989, J. Geophys. Res., 99, 18,243-18,254, 1994.

Bond, N. A, J. E. Overland, and P. Turet, Spatial and temporal characteristics of the wind forcing of the Bering Sea, J. Clim., 7, 1119-1130, 1994.

Brodeur, R. D., and N. Merati, Predation on walleye pollock (Theragra chalcogramma) eggs in the western Gulf of Alaska: the roles of vertebrate and invertebrate predators, Mar. Biol., 117, 483-493, 1993.

Brown, A. L., and K. M. Bailey, Otolith analysis of juvenile walleye pollock Theragra chalcogramma from the western Gulf of Alaska, Mar. Biol., 112, 23-30, 1992.

Buckley, L. J., RNA-DNA ratio: an index of larval fish growth in the sea, Mar. Biol., 80, 291-298, 1984.

Campana, S. E., and J. D. Neilsen, Microstructure of fish otoliths, Can. J. Fish. Aquat. Sci., 42, 1014-1032, 1985.

Canino, M. F., Effects of temperature and food availability on growth and RNA/DNA ratios of walleye pollock Theragra chalcogramma (Pallas) eggs and larvae, J. Exp. Mar. Biol. Ecol., 175, 1-16, 1994..

Canino, M. F., K. M. Bailey, and L. S. Incze, Temporal and geographic differences in feeding and nutritional condition of walleye pollock larvae Theragra chalcogramma in Shelikof Strait, Alaska, Mar. Ecol. Prog. Ser., 79, 27-36, 1991.

Clemmensen, C. M., An RNA and DNA fluorescence technique to evaluate the nutritional condition of individual marine fish larvae, Meeresforschung, 32, 134-143, 1988.

Coachman, L. K., Circulation, water masses and fluxes on the southeastern Bering Sea shelf, Cont. Shelf Res., 5, 23-108, 1986.

Davis, C. S., G. R. Flierl, P. H. Wiebe, and P. J. S. Franks, Micropatchiness, turbulence and recruitment in plankton, J. Mar. Res., 49, 109-151, 1991.

Dunn, J. R., and A. C. Matarese, A review of the early life history of Northeast Pacific gadoid fishes, Fish. Res., 5, 163-184, 1987.

Haidvogel, D. B., J. L. Wilkin, and R. Young, A semi-spectral primitive equation ocean circulation model using vertical sigma and orthogonal curvilinear horizontal coordinates, J. Comput. Phys., 94, 151-185, 1991.

Heath, M., A modeling and field study of grazing by herring larvae, Rapp. P.-v. Réun. Cons. int. Explor. Mer, 191, 233-247, 1989.

Hinckley, S., The reproductive biology of walleye pollock (Theragra chalcogramma) in the Bering Sea, with reference to spawning stock structure, Fish. Bull., 85, 481-498, 1987.

Hinckley, S., K. M. Bailey, S. J. Picquelle, J. D. Schumacher, and P. J. Stabeno, Transport, distribution, and abundance of larval and juvenile walleye pollock (Theragra chalcogramma) in the western Gulf of Alaska in 1987, Can. J. Fish. Aquat. Sci., 48, 91-98, 1991.

Hollowed, A. B., Spatial and temporal distributions of Pacific hake, Merluccius productus, larvae and estimates of survival during early life history stages, CalCOFI Rep., 33, 100-123, 1992.

Hollowed, A. B., B. A. Megrey, and W. Karp, Walleye Pollock, in Stock Assessment and Fishery Evaluation Report for the 1993 Gulf of Alaska Groundfish Fishery, prepared by the Gulf of Alaska Groundfish Plan Team for the North Pacific Fisheries Management Council, November 1992, 1993.

Houde, E. D., Fish early life dynamics and recruitment variability, Amer. Fish. Soc. Sym., 2, 17-29, 1987.

Hughes, F. W., L. K. Coachman, and K. Aagaard, Circulation, transport and water exchange in the western Bering Sea, In: Oceanography of the Bering Sea, edited by D. W. Hood and E. J. Kelley, Inst. Mar. Sci., Univ. of Alaska Occ. Pub. #2, 59-98, 1974.

Huston, M., D. DeAngelis, and W. Post, New computer models unify ecological theory, BioScience, 38, 682-691, 1988.

Incze, L. S., and T. Ainaire, Distribution and abundance of copepod nauplii and other small (40-300 m) zooplankton during spring in Shelikof Strait, Alaska, Fish. Bull., 92, 67-78, 1994.

Incze, L. S., P. B. Ortner, and J. D. Schumacher, Microzooplankton, vertical mixing and advection in a larval fish patch, J. Plankt. Res., 12, 365-379, 1990.

Incze, L. S., A. W. Kendall, Jr., J. D. Schumacher, and R. K. Reed, Interactions of a mesoscale patch of larval fish (Theragra chalcogramma) with the Alaska Coastal Current, Cont. Shelf Res., 9, 269-284, 1989.

Jones, C. M., Determining age of larval fish with the otolith increment technique, Fish. Bull. U.S., 84, 91-103, 1986.

Kendall, A. W., Jr., and S. J. Picquelle, Egg and larval distributions of walleye pollock Theragra chalcogramma in Shelikof Strait, Gulf of Alaska, Fish. Bull., U.S., 88, 133-154, 1990.

Kendall, A. W., Jr., L. S. Incze, P. B. Ortner, S. R. Cummings, and P. K. Brown, The vertical distribution of eggs and larvae of walleye pollock Theragra chalcogramma in Shelikof Strait, Alaska, Fish. Bull., 92, 540-554, 1994.

Kendall, A. W., Jr., M. E. Clarke, M. M. Yoklavich, and G. W. Boehlert, Distribution, feeding, and growth of larval walleye pollock, Theragra chalcogramma, from Shelikof Strait, Gulf of Alaska, Fish. Bull. U.S., 85, 499-521, 1987.

Kim, S., and E. P. Nunnallee, Distribution of walleye pollock, Theragra chalcogramma, spawning in Shelikof Strait, Gulf of Alaska, based on acoustic and ichthyoplankton surveys 1981, 1984 and 1985, Bull. Korean Fish. Soc., 23, 425-438, 1991.

Kim, S., and B. Bang, Oceanic dispersion of larval fish and its implication for mortality estimates: Case study of walleye pollock in Shelikof Strait, Alaska, Fish. Bull., 88, 303-311, 1990.

Kim, S., and D. R. Gunderson, Cohort dynamics of walleye pollock in Shelikof Strait, Gulf of Alaska, during the egg and larval periods, Trans. Am. Fish. Soc., 118, 264-273, 1989.

Kim, S., and A. W. Kendall, Jr., Distribution and transport of larval walleye pollock (Theragra chalcogramma) in Shelikof Strait, Gulf of Alaska, in relation to water movement, Rapp. P.-v. Réun. Cons. int. Explor. Mer., 191, 127-136, 1989.

Kope, R., and L. W. Botsford, Determination of factors affecting recruitment of chinook salmon Oncorhynchus tshawytscha in central California, Fish. Bull., 88, 257-269, 1990.

Liu, A. K., C. Y. Peng, and J. D. Schumacher, Wave-current interaction study in the Gulf of Alaska for detection of eddies by SAR, J. Geophys. Res., 99, 10,075-10,085, 1994.

McPhaden, M. J., H. B. Milburn, A. I. Nakamura, and A. J. Shepherd, PROTEUS--PROfile TElemetry of Upper ocean currentS, Sea Tech., 32, 10-19, 1991.

Macklin, S. A., P. J. Stabeno, and J. D. Schumacher, A comparison of gradient and observed over-the-water winds along a mountainous coast, J. Geophys. Res., 98, 16,555-16,569, 1993.

Megrey, B.A., Population dynamics and management of walleye pollock stocks in the Gulf of Alaska, 1976-1986, Fish. Res., 11, 321-354, 1990.

Megrey, B. A, S. J. Bograd, W. C. Rugen, A. B. Hollowed, P. J. Stabeno, S. A. Macklin, J. D. Schumacher, and W. J. Ingraham, Jr., An exploratory analysis of associations between biotic and abiotic factors and year-class strength of Gulf of Alaska walleye pollock, Can. J. Fish. Aquat. Sci. 121, 227–243, 1995.

Moore, C., J. R. V. Zaneveld, and J. C. Kitchen, Preliminary results from an in situ spectral absorption meter, Ocean Optics XI, Proc. Soc. Photo-Opt. Instrum. Eng., 1750, 330-337, 1992.

Mulligan, T. J., R. W. Chapman, and B. L. Brown, Mitochondrial DNA analysis of walleye pollock, Theragra chalcogramma, from the eastern Bering Sea and Shelikof Strait, Gulf of Alaska, Can. J. Fish. Aquat. Sci., 49, 319-326, 1992.

Mysak, L., R. D. Muench, and J. D. Schumacher, Baroclinic instability in a downstream varying channel: Shelikof Strait, Alaska, J. Phys. Oceanogr., 11, 1981.

Niebauer, H. J., Effects of El Niño-Southern Oscillation and North Pacific weather patterns on interannual variability in the subarctic Bering Sea, J. Geophys. Res., 93, 5051-5068, 1988.

Olla, B. L., and M. W. Davis, Behavioral responses of juvenile walleye pollock (Theragra chalcogramma) to light, thermoclines and food: possible role in vertical distribution, J. Exp. Mar. Biol. Ecol., in press.

Olla, B. L., and M. W. Davis, Effects of physical factors on the vertical distribution of larval walleye pollock Theragra chalcogramma under controlled laboratory conditions, Mar. Ecol. Prog. Ser., 63, 105-112, 1990.

Overland, J. E., M. C. Spillane, H. E. Hurlburt, and A. J. Wallcraft, A numerical study of the circulation of the Bering Sea basin and exchange with the North Pacific Ocean, J. Phys. Oceanogr., 24, 736-758, 1994.

Perry, R. I., Book Reviews, Fish. Oceangr., 3, 218-220, 1994.

Picquelle, S. J., and B. A. Megrey, A preliminary spawning biomass estimate of walleye pollock Theragra chalcogramma in the Shelikof Strait, Alaska, based on the annual egg production method, Bull. Mar. Sci., 53, 728-749, 1993.

Reed, R. K., and S. J. Bograd, Transport in Shelikof Strait, Alaska: An update, Cont. Shelf Res. 15(2/3), 213-218, 1995.

Reed, R. K., and P. J. Stabeno, Flow along and across the Aleutian Ridge, J. Mar. Res., 52, 639-648, 1994.

Reed, R. K., and P. J. Stabeno, The return of the Alaskan Stream to Near Strait, J. Mar. Res., 51, 515-527, 1993.

Reed, R. K., and J. D. Schumacher, Sea level variations in relation to coastal flow around the Gulf of Alaska, J. Geophys. Res., 86, 6543-6546, 1981.

Reed R. K., and J. D. Schumacher, Transport and physical properties in central Shelikof Strait, Alaska, Cont. Shelf Res., 9, 261-268, 1989a.

Reed R. K., and J. D. Schumacher, Some mesoscale features of flow in Shelikof Strait, Alaska, J. Geophys. Res., 94, 12,603-12,606, 1989b.

Reed, R. K., J. D. Schumacher, and L. S. Incze, Circulation in Shelikof Strait, Alaska, J. Phys. Oceanogr., 17, 1546-1554, 1987.

Reed, R. K., L. S. Incze, and J. D. Schumacher, Estimation of the effects of flow on dispersion of larval pollock Theragra chalcogramma in Shelikof Strait, Alaska, in Effects of Ocean Variability on Recruitment and an Evaluation of Parameters Used in Stock Assessment Models, edited by R. J. Beamish and G. A. McFarlane, Can. J. Fish. Aquat. Sci., Spec. Pub., 108, pp. 239-246, 1989.

Reed, R. K., J. D. Schumacher, and A. W. Kendall, Jr., NOAA's Fisheries Oceanography Coordinated Investigations in the western Gulf of Alaska, Eos, Trans. AGU, 69, 890-894, 1988.

Reed, R. K., J. D. Schumacher, P. J. Stabeno, C. H. Pease, A. W. Kendall, Jr., and P. Dell'Arciprete, Alaskan Stream inflow and effects in the Bering Sea, Eos, Trans. AGU, 75, 342, 1994.

Reed, R. K., G. V. Khen, P. J. Stabeno, and A. V. Verkhunov, Water properties and flow over the deep Bering Sea basin, summer 1991, Deep-Sea Res., 40, 2325-2334, 1993.

Rothschild, B. J., Dynamics of marine fish populations, Harvard University Press, Cambridge, MA, 277 pp., 1986.

Royer, T. C., Coastal fresh water discharge in the northeast Pacific, J. Geophys. Res., 87, 2017-2021, 1982.

Royer, T. C., Baroclinic transport in the Gulf of Alaska, Part II, Fresh water driven coastal current, J. Mar. Res., 39, 251-266, 1981.

Schumacher, J. D., and P. J. Stabeno, Ubiquitous eddies of the Eastern Bering Sea and their coincidence with concentrations of larval pollock, Fish. Oceanogr., 3, 182-190, 1994.

Schumacher, J. D., and R. K. Reed, Characteristics of currents over the continental slope of the eastern Bering Sea, J. Geophys. Oceanogr., 97, 9423-9433, 1992.

Schumacher, J. D., and A. W. Kendall, Jr., Some interactions between young walleye pollock and their environment in the western Gulf of Alaska, CalCOFI Rep., 32, 22-40, 1991.

Schumacher, J. D., and T. H. Kinder, Low-frequency current regimes over the Bering Sea shelf, J. Phys. Oceanogr., 13, 607-623, 1983.

Schumacher, J. D., and R. K. Reed, Coastal flow in the northwest Gulf of Alaska: the Kenai Current, J. Geophys. Res., 85, 6680-6688, 1980.

Schumacher, J. D., P. J. Stabeno, and S. J. Bograd, Characteristics of an eddy over a continental shelf: Shelikof Strait, Alaska, J. Geophys. Res., 98, 8395-8404, 1993.

Schumacher, J. D., P. J. Stabeno, and A. T. Roach, Volume transport in the Alaska Coastal Current, Cont. Shelf Res., 9, 1071-1089, 1989.

Schumacher, J. D., W. E. Barber, B. Holt, and A. K. Leu, Satellite observations of mesoscale features in Lower Cook Inlet and Shelikof Strait, Alaska, NOAA Tech. Rep. ERL-445-PMEL 40 (NTIS# PB92-101633), 18 pp., 1991.

Severin, K. P., J. Carroll, and B. L. Norcross, Electron microprobe analysis of juvenile walleye pollock (Theragra chalcogramma) otoliths from Alaska: a pilot stock separation study, Env. Biol. Fish., in press.

Smith, S. L., and J. Vidal, Variations in the distribution, abundance, and development of copepods in the southeastern Bering Sea in 1980 and 1981, Cont. Shelf. Res., 5, 215-239, 1986.

Springer, A. M., A review: Walleye pollock in the North Pacific--how much difference do they really make?, Fish. Oceanogr., 1, 80-96, 1992.

Stabeno, P. J., and R. K. Reed, Circulation in the Bering Sea basin by satellite tracked drifters, J. Phys. Oceanogr., 24, 848-854, 1994.

Stabeno, P. J., and R. K. Reed, A major circulation anomaly in the western Bering Sea, Geophys. Res. Letters, 19, 1671-1674, 1992.

Stabeno, P. J., and R. K. Reed, Recent Lagrangian measurements along the Alaskan Stream, Deep-Sea Res., 38, 289-296, 1991.

Stabeno, P. J., R. K. Reed, and J. D. Schumacher, The Alaska coastal current: continuity of transport and forcing, J. Geophys. Res. 100(C2), 2477-2485, 1995.

Stabeno, P. J., A. J. Hermann, N. A. Bond, and S. J. Bograd, Modeling the impact of climate variability on the advection of larval walleye pollock (Theragra chalcogramma) in the Gulf of Alaska, Can. J. Fish. and Aquat. Sci., 121, 719-727, 1995.

Theilacker, G. H., Effects of starvation on the historical and morphological characteristics of jack mackerel, Trachurus symmerticus, larvae, Fish. Bull. U. S., 76, 403-414, 1978.

Theilacker G. H., and W. Shen, Calibrating starvation-induced stress in larval fish using flow cytometry, Amer. Fish. Soc. Sympos., 14, 85-94, 1993.

Theilacker G. H., N. C. H. Lo, and A. W. Townsend, An immunochemical approach to quantifying predation by euphausiids on the early life stages of anchovy, Mar. Ecol. Prog. Ser., 92, 35-50, 1993.

Tyler, A. V., A context for recruitment correlations: why marine fisheries biologists should still look for them, Fish. Oceanogr., 1, 97-107, 1992.

Vastano, A. C., L. S. Incze, and J. D. Schumacher, Observation and analysis of fishery processes: larval pollock at Shelikof Strait, Alaska, Fish. Oceanogr., 1, 20-31, 1992.

Walters, C. J., and J. S. Collie, Is research on environmental factors useful for fisheries management?, Can. J. Fish. Aquat. Sci., 45, 1848-1854, 1988.

Westpested, V. G., The status of Bering Sea pollock and the effect of the "Donut Hole" Fishery, Fisheries, 18, 18-24, 1993.

Wooster, W. S., Immiscible Investigators: Oceanographers, Meteorologists, and Fishery Scientists, in Natural Resources, Economic and Policy Applications, edited by E. Miles, R. Pearly, and R. Stokes, University of Washington Press, Seattle, WA, pp. 374-386, 1986.

Yoklavich, M. M., and K. M. Bailey, Hatching period, growth and survival of young walleye pollock Theragra chalcogramma as determined from otolith analysis, Mar. Ecol. Prog. Ser., 64, 13-2, 1990.


Return to Abstract

PMEL Outstanding Papers

PMEL Publications Search

PMEL Homepage