GLERL Publications Abstracts: FY 1983

ASSEL, R.A. A computerized ice concentration data base for the Great Lakes. NOAA Data Report ERL GLERL-24, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-233031) 28 pp. (1983). https://www.glerl.noaa.gov/pubs/tech_reports/glerl-024/dr-024.pdf

A 20-winter computerized ice concentration data base was established for the great Lakes. This report describes the computerized data set and an ice concentration climatology developed from it. Data reduction and analysis procedures, computer file structure and record format, and availability of the data are given.

ASSEL, R.A. Description and analysis of a 20-year (1960-79) digital ice-concentration database for the Great Lakes of North America. Annals of Glaciology 4:14-18 (1983).

A digital ice-concentration database spanning 20 years (1960 to 1979) was established for the Great Lakes of North America. Data on ice concentration, i.e. the percentage of a unit surface area of the lake that is ice-covered, were abstracted from over 2 800 historic ice charts produced by United States and Canadian government agencies. The database consists of ice concentrations ranging from zero to 100% in 10% increments for individual grid cells of size 5 x 5 km constituting the surface area of each Great Lake. The data set for each of the Great Lakes was divided into half-month periods for statistical analysis. Maximum, minimum, median, mode, and average ice-concentrations statistics were calculated for each grid cell and half-month period. A lakewide average value was then calculated for each of the half-month ice-concentration statistics for all grid cells for a given lake. Ice-cover variability and the normal extent and progression of the ice cover is discussed within the context of the lakewide averaged value of the minimum and maximum ice concentrations and the lakewide averaged value of the median ice concentrations, respectively. Differences in ice-cover variability among the five Great Lakes are related to mean lake depth and accumulated freezing degree-days. A Great Lakes ice atlas presenting a series of ice charts which depict the maximum, minimum, and median icecover concentrations for each of the Great Lakes for nine half-monthly periods, starting the last half of December and continuing through the last half of April will be published in 1933 by the National Oceanic and Atmospheric Administration (NOAA). The database will be archived at the National Snow and Ice Data Center of the National Environmental Satellite Data and Information Service (NESDIS) in Boulder, Colorado, USA, also in 1983.

ASSEL, R.A. Lake Superior bathythermograph data: 1973-79. NOAA Data Report ERL GLERL-25, Great Lakes Environmental Research Laboratory (PB83-252890) 24 pp. (1983).

Temperature surveys were made across an east-west transect of Lake Superior during the 1973, 1974, 1975, and 1976 winter seasons and the 1976, 1977, 1978, and 1979 autumn seasons. A portable bathythermograph recorder measured water temperatures along the transect to a maximum depth of 200 m. There were 46 temperature surveys made, with an average of 24 temperature profile measurements per survey. This report presents tabulations of 39 of the 46 temperature surveys, along with isothermal contour charts portraying the thermal structure of the lake and station position charts showing the location of each temperature profile for all 46 surveys. Temperature tabulations from the 1973 and 1974 winters were published in a previous report.

BENNETT, J.R., A.H. CLITES, and D.J. SCHWAB. A two-dimensional lake circulation modeling system: Programs to compute particle trajectories and the motions of dissolved substances. NOAA Technical Memorandum ERL GLERL-46, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-257014) 58 pp. (1983). https://www.glerl.noaa.gov/pubs/tech_reports/glerl-046/

This report documents two computer programs that use currents from numerical lake circulation models to predict the motion of particles and dissolved substances. Movement of particles in the water is related to the vertically averaged current and to a percentage of the surface wind. Dissolved substances are assumed to be vertically mixed and thus to move only with the current. The user must provide bathymetric data, initial conditions, currents, winds, and a subroutine to generate output. The programs are very general and, though designed to be used with the hydrodynamic models and bathymetric grid generation programs documented in earlier GLERL technical memoranda, can be easily adapted to use input from other sources.

Bowling, J.W., G.J. Leversee, P.F. LANDRUM, and J.P. Giesy. Acute mortality of anthracene-contaminated fish exposed to sunlight. Aquatic Toxicology 3:79-90 (1983).

Acute mortality of bluegill sunfish (Lepomis macrochirus) dosed with anthracene at 12.7 mg/l and exposed to natural sunlight conditions was observed during a study of anthracene fate in outdoor channel microcosms. No mortality was observed under control conditions (natural sunlight and no anthracene). Fish survived when held in the shade downstream of sunlit contaminated water, arguing against mortality due to toxic anthracene photoproducts in the water. Fish held 48 h in anthracene contaminated water (- 12 mg/1), in a shaded channel, died when placed in clean water and exposed to sunlight. After 144 h depuration in darkness, fish anthracene concentrations had decreased to pre-exposure concentrations and no mortality was observed when fish were subsequently exposed to sunlight. This observed photoinduced toxic response in anthracene contaminated fish may represent a significant environmental hazard of polycyclic aromatic hydrocarbons in aquatic environments.

CROLEY, T.E.II. A tank-cascade runoff model for large forested basins. Proceedings, Canadian Hydrology Symposium '82: Hydrological Processes of Forested Areas, Ottawa, ON, Canada, June 14-15, 1982. Associate Committee on Hydrology, National Research Council Canada, Fredericton, New Brunswick, 419-44 (1982).

Large-scale watershed models are required to estimate basin runoff to the Great Lakes for use in routing determinations and operational hydrology studies. Data limitations, large-basin applicability, and economic efficiency preclude the use of existing large-watershed models. This paper describes an interdependent tank-cascade model that uses a mass balance coupled with linear reservoir concepts. It is physically based and uses climatological considerations not possible for small watersheds; it employs analytical solutions to bypass numerical inaccuracies. Snowmelt and net supply computations are separable from the mass balance determinations and are based on a simple heat balance. Partial area concepts are used to determine infiltration and surface runoff. Losses are determined from joint consideration of available energy for actual and potential evapotranspiration and of available moisture in the soil horizons by using climatological concepts, especially relevant for large data-poor forested areas. The model is applied, for a 30-day computation interval, to the Genesee River Basin in New York State and compared with past 6-hr computation interval applications of the Streamflow Synthesis and Reservoir Regulation and National Weather Service Hydrologic models to the same data set. Results are summarized also for 11 other large watersheds about Lake Ontario.

CROLEY, T.E.II. Great Lakes basins (U.S.A.-Canada) runoff modeling. Journal of Hydrology 64:135-158 (1983).

Large-scale watershed models are required in order to estimate basin runoff to the Great Lakes for use in routing determinations and operational hydrology studies. Data limitations, large-basin applicability and economic efficiency preclude the use of existing large-watershed models. This paper describes an interdependent tank-cascade model that uses a mass balance coupled with linear reservoir concepts. It is physically based and uses climatological considerations not possible for small watersheds; it employs analytical solutions to bypass numerical inaccuracies. Snowmelt and net-supply computations are separable from the mass-balance determinations and are based on a simple heat balance. Partial-area concepts are used to determine infiltration and surface runoff. Losses are determined from joint consideration of available energy for evapotranspiration and of available moisture in the soil horizons by using climatologic concepts. Also described are heuristic calibration procedures that give insight into the use of the model. The model is applied, for a 30-day computation interval, to the Genesee River Basin in New York State and compared with past 6-hr computation interval applications of the Streamflow Synthesis and Reservoir Regulation (SSARR) and National Weather Service Hydrologic (NWSH) models to the same data set.

CROLEY, T.E.II. Lake Superior Basin runoff modeling. GLERL Open File Report, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (1983).

No abstract.

CROLEY, T.E.II. Sediment dynamics in unsteady nonprismatic rills. Proceedings, Conference on Frontiers in Hydraulic Engineering, H.T. Shen (ed.), Massachusets Institute of Technology, Cambridge, MA, August 9-12, 1983. American Society of Civil Engineers, New York, 127-132 (1983).

Overland sedimentation in unsteady nonprismatic rills under uniform rainfall excess appears to be well described with kinematic models of flow, entrainment, and deposition applied to developing flow geometries; this is not possible with sheet-flow models. Sediment concentration and flow data are analyzed by extending the theory developed for prismatic channels. Excellent agreement between theory and experimental results at several rainfall rates is demonstrated here for one high rainfall rate. The rills proceed (in both time and space) from narrow and (relatively) deep channels (early or upstream) to wide and shallow (later or downstream), resulting in a decrease of average velocity with increasing flow thereby increasing deposition. A reversal of sediment and flow characteristic times is apparent, compared with prismatic channel analyses, since the flow velocity, v, decreases with increasing flow area, A (dv/da < 0).

CROLEY, T.E., II, and H.C. HARTMANN. Lake Ontario Basin runoff modeling. NOAA Technical Memorandum ERL GLERL-43, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-237420) 117 pp. (1983).

An interdependent tank-cascade model of basin runoff, employing analytical solutions of climatological considerations relevant for large watersheds, has been developed. The mass balance is coupled with physically-based concepts of linear reservoir storages, partial-area infiltration, complementary evapotranspiration and evapotranspiration opportunity based on available supply, and heat balance determinations of snowmelt and net supply. Daily air temperature, precipitation, and runoff data are required for calibration of the nine parameters; data are grouped for 15 watersheds about Lake Ontario, as well as for the entire basin above Elevator, NY, and Kingston, Ontario. The model has been applied to the Lake Ontario Basin in both lumped- and distributed-parameter approaches; 11 subbasins and 2 basins have been modeled for 7-d and 30-d mass balance computation intervals. Parameter values have been interpreted for physical meaning and relation to data errors and computation intervals. Temporal and spatial integration effects have been analyzed with respect to error reduction, modeling information and resolution, and cost trade-offs. The model is an accurate, fast representation of weekly or monthly runoff volumes from large watersheds with simple calibration and data requirements. Parameter values have physical significance and appear reasonable and consistent.

DERECKI, J.A. Travel times in the Great Lakes connecting channels. GLERL Open File Report, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (1983). https://www.glerl.noaa.gov/pubs/fulltext/1983/19830006.pdf

No abstract.

EADIE, B.J., W.R. FAUST, P.F. LANDRUM, N.R. MOREHEAD, W.S. GARDNER, and T.F. NALEPA. Bioconcentrations of PAH by some benthic organisms. Proceedings, Polynuclear Aromatic Hydrocarbons: Seventh International Symposium on Formation, Metabolism, and Measurement, M.W. Cooke and A.J. Dennis (eds.), Battelle Press, Columbus, OH, 437-449 (1982).

No abstract.

EADIE, B.J., W.R. FAUST, P.F. LANDRUM, and N.R. MOREHEAD. Factors affecting bioconcentration of PAH by the dominant benthic organisms of the Great Lakes. Proceedings, Polynuclear Aromatic Hydrocarbons: Eighth International Symposium on Mechanisms, Methods, and Metabolism, M.W. Cooke and A.J. Dennis (eds.), Battelle Press, Columbus, OH, 799-812 (1983).

No abstract.

EADIE, B.J., P.F. LANDRUM, and W.R. FAUST. Polycyclic aromatic hydrocarbons in sediments, pore water, and the amphipod Pontoporeia hoyi from Lake Michigan. Chemosphere 11:847-858 (1982).

Polycyclic aromatic hydrocarbon (PAH) concentrations were measured in Pontoporeia hoyi, the most abundant benthic organism in Lake Michigan, and in its associated sediment/pore water matrix. Individual PAH concentrations ranged from 10 ppb to 1 ppm in three sedimentary environments having different levels of organic carbon. Pore water concentrations appeared to be independent of sediment concentrations. P. hoyi bioconcentration factors ranged from approximately 104 to 105 for seven analyzed PAHs. A major fraction of the phenanthrene, fluoranthene and pyrene in P. hoyi appears to come from sediments and pore water, while chrysene and BaP are primarily obtained from water.

EADIE, B.J., C.P. Rice, and W.A. FREZ. The role of the benthic boundary layer in the cycling of PCB's in the Great Lakes. In Physical behavior of PCB's in the Great Lakes, D. MacKay (ed.). Ann Arbor Science, Ann Arbor, MI, 213-229 (1982).

No abstract.

EADIE, B.J., J.A. ROBBINS, P.F. LANDRUM, C.P. Rice, M.S. Simmons, M.J. McCORMICK, S.J. Eisenreich, G.L. BELL, R.L. PICKETT, K. Johansen, K. Rossman, N. HAWLEY, and T. Voice. The cycling of toxic organics in the Great Lakes: A 3-year status report. NOAA Technical Memorandum ERL GLERL-45, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-256792) 176 pp. (1983). https://www.glerl.noaa.gov/pubs/tech_reports/glerl-045/

This interim 3-year status report describes the results of GLERL's studies on the cycling of toxic organics in the Great Lakes. A hierarchy of models has been developed including 1) a lakescale equilibrium model, 2) a one-dimensional steady-state model, 3) one and two-dimensional time dependent models, and 4) several individual process models. These modeling efforts have identified process research needs, some of which have been supported. Reported here are results of our work on: 1) air-water exchange; 2) photolysis; 3) sorption and partitioning; 4) particle settling and transport; 5) early diagenetic processes in lake sediments; 6) interaction of sediments, contaminants, and benthic organisms; and 7) simulation studies of organic contaminants.

Edgington, D.N., and J.A. ROBBINS. Patterns of deposition of natural and fallout radionuclides in the sediments of Lake Michigan and their relation to limnological processes. In Environmental Biochemistry, J.O. Nriagu (ed.). Ann Arbor Science, Ann Arbor, MI, 705-729 (1983).

No abstract.

GARDNER, W.S., and P.F. LANDRUM. Characterization of ambient levels of ultraviolet-absorbing dissolved humic materials in natural waters by aqueous liquid chromatography. In Aquatic and Terrestrial Humic Materials, R.F. Christman and E.T. Gjessing (eds.). Ann Arbor Science, Ann Arbor, MI, 203-217 (1983).

Ambient levels of ultraviolet-absorbing dissolved organic materials of river, lake and sediment pore water filtrates were fractionated into chemically distinct groups by high-performance, distilled water, size exclusion chromatography. When the mobile phase was distilled water, fractionation apparently depended on compound polarity, as well as on molecular weight. Water samples (0.5 ml) were analyzed directly on a TSK 3000 sw size exclusion column with ultraviolet (UV) detection. The chemical character of the three (or four) resolved peaks was examined by pretreatments, including ultra centrifugation; passage through cation exchange, anion exchange, and reverse phase columns; metal addition; and dialysis. The first two peaks could be removed by anion exchange (but not by reverse phase or cation exchange) pretreatment. Components of the third peak were not quantitatively removed by anion exchange, cation exchange, or reverse phase pretreatments. Peak 4, found only in sediment pore water, was removed by reverse phase pretreatment but not by anion or cation exchange. The addition of Na, Ca or Cu to filtrates resulted in removal of peaks 1 and 2, with a concomitant increase in peak 3. Peak 1 was removed progressively less efficiently by Cu, Ca and Na, but in contrast, peak 2 was removed by similar levels of each respective metal. Thus, when distilled water was used as the mobile phase, components of each UV-absorbing peak exhibited unique chemical characteristics.

GARDNER, W.S., and J.M. MALCZYK. Discrete injection segmented flow analysis of nutrients in small-volume water samples. Analytical Chemistry 55:1645-1647 (1983).

No abstract.

GARDNER, W.S., T.F. NALEPA, D.R. SLAVENS, and G.A. LAIRD. Patterns and rates of nitrogen release by benthic Chironomidae and Oligochaeta. Canadian Journal of Fisheries and Aquatic Sciences 40(3):259-266 (1983).

Metabolic mineralization of ammonium by tubificid worms and chironomid larvae appears to be an important mechanism contributing to nitrogen regeneration from aerobic lake sediments. Mean weight [ash-free dry weight (AFDW)] - specific ammonium release rates ranged from 3 to 15 nmol NH4 (mg AFDW)-1 x h-1 for chironomid and tubificid species collected at different times and temperatures from nearshore Lake Michigan sediments. Although mean rates of nitrogen release were similar for the two groups of benthic invertebrates, the patterns of release were different. Tubificids released nitrogen (ammonium plus primary amines) continuously, whereas chironomids released it in spurts several times per hour. Mean ammonium-release rates were generally constant with time after the animals were removed from food for both species. This implies that ammonium regeneration for these benthic animals is primarily an endogenous process. Conversely, primary amine nitrogen release often decreased with time after food removal and may reflect either egestion of partially digested materials from the animals' guts or a decrease in the metabolic pool size of free amino acids. Based on previous measurements of phosphorus release, mean molar NH4:PO4 ratios in excreted materials were calculated to be 15:1 for chironomids and 35:1 for tubificids.

GARDNER, W.S., and H.A. VANDERPLOEG. Microsample-filtering device for liquid chromatography or flow injection analysis. Analytical Chemistry 54:2129-2130 (1982).

No abstract.

Gerould, S., P.F. LANDRUM, and J.P. Giesy. Anthracene bioconcentration and biotransformation in chironomids: Effects of temperature and concentration. Environmental Pollution (Series A) 30:175-188 (1983).

Effects of temperature and anthracene concentration on uptake (Ku) and depuration (Kd) rate constants and bioconcentration factor (14C-BCF) were determined for larvae of the midge Chironomus riparius. At constant temperature (25oC) the uptake rate constant estimated from 10 h and 30 h exposure and by the initial rates methods increased with concentration between 1 x 7 and 30 x 5 mg litre-1. At constant concentration (22mg litre-1), the uptake rate constant was maximum at 25oC and less at 16o and 30oC. The apparent increase in depuration rate constant with concentration during 30 h exposure was not confirmed in experiments in which contaminated animals depurated in uncontaminated paper towel. The 14C-BCF did not change as a function of temperature or anthracene concentration. BCF based on anthracene concentration was minimum at 25oC when biotransformation rate was highest, and was more than an order of magnitude lower than 14C-BCF.

GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY. Annual Report for the Great Lakes Environmental Research Laboratory, FY 1982. Great Lakes Environmental Research Laboratory, Ann Arbor, MI, 45 pp. (1982).

No abstract.

GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY. Detailed technical plan for the Great Lakes Environmental Research Laboratory. Great Lakes Environmental Research Laboratory, Ann Arbor, MI, 252 pp. (1983).

No abstract.

GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY. Technical plan for the Great Lakes Environmental Research Laboratory. Great Lakes Environmental Research Laboratory, Ann Arbor, MI, 68 pp. (1983).

No abstract.

*GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY, The University of Michigan, The University of Minnesota, Michigan State University, Argonne National Laboratory, and Oak Ridge National Laboratory. Annual Report to the Ocean Assessment Division/National Ocean Service, NOAA. In The Cycling of Toxic Organic Substances in the Great Lakes Ecosystem, Great Lakes Environmental Research Laboratory, Ann Arbor, MI, (1983).

No abstract.

*GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY, The University of Michigan, and The University of Minnesota. Fourth semiannual progress report to the Office of Marine Pollution Assessment, NOAA. In The Cycling of Toxic Organic Substances in the Great Lakes Ecosystem, Great Lakes Environmental Research Laboratory, Ann Arbor, MI, 61 pp. (1983).

No abstract.

*GREAT LAKES ENVIRONMENTAL RESEARCH LABORATORY, The University of Michigan, The University of Minnesota, Michigan State University, Argonne National Laboratory, and Oak ridge National Laboratory. Semiannual Report to the Office of Marine Pollution Assessment, NOAA. In The Cycling of Toxic Organic Substances in the Great Lakes Ecosystem, Great Lakes Environmental Research Laboratory, Ann Arbor, MI, (1983).

No abstract.

GREENE, G.M. Forecasting ice-cover freeze-up, growth, and breakup on the St. Marys River. NOAA Technical Memorandum ERL GLERL-47, Great Lakes Environmental Research Laboratory (PB83-262097) 82 pp. (1983).

A 10-year time series of meteorological variables, water temperatures, and ice observations was used to develop methods for the prediction of ice-cover formation, growth rates, and decay at five sites along the St. Marys River, the channel connecting Lake Superior and Lake Huron. A site-specific heat transfer coefficient and observed water temperatures at Sault Ste. Marie, Mich., can be used to predict ice-cover formation. Standard errors in the predictions at the five sites are 30- to 60-percent lower than the corresponding standard deviations of the observations. A simple Stefan relationship with an average standard error of 8 cm over the season can be used to simulate ice-cover growth. Unlike the ice formation prediction method, ice growth prediction is quite sensitive to the accuracy of the air temperature forecasts. No one method can be used to predict ice-cover breakup at all five sites. Breakup dates are most strongly correlated with the date at which water temperature rises above OoC at Sault Ste. Marie. This date, however, can be less than 1 week prior to breakup at some sites or may occur after breakup. Maximum ice-cover thickness in the river and maximum ice-cover extent on Lake Superior are both poor predictors of the breakup date.

GREENE, G.M., and F. Nelson. Performance of a frost hollow as a hemispherical thermal radiometer. Archives for Meteorology, Geophysics, and Bioclimatology, Series B 32:263-278 (1983).

No abstract.

GUERRA, B.E., and W.C. Sonzogni. Wind erosion as a source of water pollution--A brief update. GLERL Open File Report, Great Lakes Environmental Research Laboratory, Ann Arbor, MI, 2 pp. (1983).

No abstract.

Haddock, J.D., P.F. LANDRUM, and J.P. Giesy. Extraction efficiency of anthracene from sediments. Analytical Chemistry 55:1197-1200 (1983).

No abstract.

HAWLEY, N. A numerical model of cohesive suspended sediment dynamics. NOAA Technical Memorandum ERL GLERL-42, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-207860) 41 pp. (1983). https://www.glerl.noaa.gov/pubs/tech_reports/glerl-042/

This report documents a one-dimensional finite-difference computer program that models cohesive suspended sediment dynamics in a shear flow. The model is based on Smulchowski's geometrical collision formulas. User-supplied empirical constants are necessary to determine the collision efficiency and aggregate shear strength. The model does not include biological or chemical processes or lateral advection. At present, the model is designed to reflect conditions in the Great Lakes, but by changing the boundary conditions it could be modified for other environments.

HAWLEY, N. Intertidal sedimentary structures on macrotidal beaches. Journal of Sedimentary Petrology 52(3):785-796 (1983).

A one-month survey of three macrotidal beaches located in south Wales has shown that the distribution of bed forms in the intertidal zone is controlled primarily by wave climate and tidal range. The areas near the high- and low-water marks reflect the predominance of swash-zone action, whereas the middle part of the beach is more likely to contain bed forms formed in either the breaking or shoaling wave zones. Since none of the structures were formed by current action, changes in tidal range affected only the width of the areas, not the type of bed forms produced. The bed forms observed varied directly with changes in wave height; on high wave-energy beaches, the model of Clifton et al. (1971) is valid, whereas on lower energy beaches it must be modified. the rapid response of the beaches to changes in wave climate makes it unlikely that be forms will be preserved except possibly during extremely severe storms, when the lowermost part of the be forms might be preserved as part of a vertical sequence produced by migrating facies.

HAWLEY, N. Lake Michigan suspended sediment characteristics at Grand Haven, Michigan, 1979. NOAA Data Report ERL GLERL-23, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83222638) 3 pp. (1983).

Suspended sediments collected on 4 different days in 1979 have been measured with a Quantimet image analyzer and identified by particle type. The data may be sued to deduce changes in particle composition as a function of particle size, season of the year, and water depth.

HAWLEY, N. Reply. Sedimentology 29:906-907 (1982).

No abstract.

HAWLEY, N., R.L. CHAMBERS, and G.L. BELL. Grain-size distribution in Lake Michigan, 1977-81. NOAA Data Report ERL GLERL-21, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-194399) 9 pp. (1983).

The grain-size distributions of suspended material collected from Lake Michigan were measured with a HIAC particle counter. The Lake Michigan samples appear to have relatively more large particles than do oceanic samples, but the difference may be due to differences in the instruments used to measure the particles.

Henderson-Sellers, B., M.J. MCCORMICK, and D. SCAVIA. A comparison of the formulation for eddy diffusion in two one-dimensional stratification models. Applied Mathematical Modelling 7:212-215 (1983).

Two models for thermal stratification based on turbulent diffusion concepts are analyzed and compared. The models by Henderson-Sellers; and by McCormick and Scavia, are shown to be equivalent at large values of the Richardson number, Rj. At small Rj, the simpler model reverts to specification of the turbulent diffusion as a constant value. This simplification is also demonstrated to be a realistic approximation only at low wind speeds and for deep lakes. By comparison of these model types, a (previously empirically defined by McCormick and Scavia) parameter b is related conceptually to the lake depth, H.

HINKEL, K.M. Ice-cover growth rates at nearshore locations in the Great Lakes. NOAA Technical Memorandum ERL GLERL-44, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-233049) 35 pp. (1983). https://www.glerl.noaa.gov/pubs/tech_reports/glerl-044/

Ice thickness data from 32 nearshore locations around the Great Lakes were correlated to accumulated degree-days of frost over a 6- to 11-year period. A simple parastatistical model was used to compute ice-cover growth coefficients that reflect the relative impact of site-specific factors and processes on ice growth for each ice measurement site. In addition, two data sets were used to illustrate the inhibiting influence of snow on ice growth. Statistical parameters generated for each site and data set were used to summarize the degree of predictive accuracy. For 27 sites, a weighted R2 value of 0.82 was achieved with an average standard error of estimate of 6.95 cm. As an additional test, the site-specific ice growth coefficients were applied to unpublished ice thickness data for the abnormally cold winters of 1977-78 and 1978-79. For these two seasons, the average standard error of estimate was 5.39 cm.

International Joint Commission, InternationalGreatLakes Technical Information Network Board. Great Lakes hydrometeorological station directory. NOAA Data Report ERL GLERL-22, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-208124) 259 pp. (1983).

This report lists all hydrological stations located within the Great Lakes Basin and all meteorological stations located in or near the Great Lakes Basin deemed pertinent to the International Great Lakes Technical Information Network Board. The listing identifies each station by station number, station name, location, elevation, years of record, type, and availability of the data, operating agency, and acquisition agency. A standard table format was created and used throughout. All information presented was provided by various agencies within the United States and Canada. Stations were grouped by type and plotted to provide information on the spatial distribution of the various types of stations. The base map is of the entire Great Lakes and St. Lawrence River drainage basins and includes each Great Lakes subbasin.

LANDRUM, P.F. The effect of co-contaminants on the bioavailability of polycyclic aromatic hydrocarbons to Pontoporeia hoyi. Proceedings, Polynuclear Aromatic Hydrocarbons: Seventh International Symposium on Formation, Metabolism, and Measurement, M.W. Cooke and A.J. Dennis (eds.), Battelle Press, Columbus, OH, 731-743 (1982).

No abstract.

LANDRUM, P.F. Uptake, depuration, and biotransformation of anthracene by the scud Pontoporeia hoyi. Chemosphere 11(10):1049-1057 (1982).

Uptake, depuration and biotransformation rates of 14C-anthracene were determined for Pontoporeia hoyi, the dominant benthic invertebrate in the Great Lakes, at 4o, 7o, 10o and 15oC. The uptake rate constants for anthracene increased from 136+/-22 h-1 (n=4, X+/-1SD) to 215+/-45 h-1 (n=4) over the temperature range studied and were seasonally dependent. The depuration rate constant at the apparent optimum temperature of 7oC was 0.015 h-1 for anthracene. The biotransformation ability of P. hoyi is low, and degradation of anthracene was undetectable even after exposures of 48 h. The bioconcentration factor can be predicted from the uptake and depuration kinetics to be approximately 16,800 at 4oC. These experiments imply that P. hoyi may be very important in food chain biomagnification of some toxic organics.

LANDRUM, P.F., B.J. EADIE, W.R. FAUST, N.R. MOREHEAD, and M.J. McCORMICK. Role of sediment in the bioaccumulation of benzo(a)pyrene by the amphipod Pontoporeia hoyi. Proceedings, Polynuclear Aromatic Hydrocarbons: Eighth International Symposium on Mechanisms, Methods, and Metabolism, M.W. Cooke and A.J. Dennis (eds.), Battelle Press, Columbus, OH, 799-812 (1983).

No abstract.

LANDRUM, P.F., and D. SCAVIA. Influence of sediment on anthracene uptake, depuration, and biotransformation by the amphipod Hyalella azteca. Canadian Journal of Fisheries and Aquatic Sciences 40(3):298-305 (1983).

Uptake, depuration, and biotransformation rates of anthracene were determined for Hyalella azteca (Amphipoda, Crustacea) in both the presence and the absence of sediment. The mean rate constant +/-se for uptake from water was the same for cases with or without sediment, 255 +/- 76 mL x (g animal wet weight)-1 x h-1. The rate constant for uptake of sediment-associated athracene was 19 +/- 5 g dry sediment x (g animal wet weight)-1 x h-1 for an organic sediment. Depuration increased by a factor of three in the presence of sediment. Sediment-associated anthracene, compound sorbed to sediment and in pore water, was estimated to contribute 77% of the steady-state body burden. These animals, and presumably other benthic organisms exposed to toxicants in organic sediments, are at higher risk than animals not associated with sediments. Further, bioconcentration factors based on the ratio of body burden to water concentration are overestimates for H. azteca and other benthic organisms that obtain a large proportion of their body burden from sediment.

LIU, P.C. An iteration problem: Problem 83-9. SIAM Rev. 25:268 (1983).

No abstract.

McCORMICK, M.J., B.L. WHARRAM, and D. SCAVIA. Comparison of phytoplankton growth constructs in the context of vertically segmented plankton models. In Developments in Environmental Modelling, Part 5. Analysis of Ecological Systems: State-of-the-art in Ecological Modeling, W.K. Lauenroth, G. V. Skogerboe, and M. Flug (eds.). Elsevier Science Publishers, Amsterdam, The Netherlands, 663-668 (1982).

A previously developed and calibrated ecological model, extended to a finer vertical grid representation, was capable of simulating the vertical structure of Lake Ontario's phytoplankton, available nitrogen and silicon, but not available phosphorus. Measured phosphorus depletion extended to 40 m whereas nitrogen and silicon depletion extended to only 5-10 m. Analysis of models using different algal growth constructs demonstrated the ability of algal models based upon internal nutrient pools to simulate the observed divergence between phosphorus and other nutrient profiles.

Meadows, G.A., N. Schultz, H. Dannelongue, J.R. BENNETT, J.E. CAMPBELL, P.C. LIU, and D.J. SCHWAB. The response of the coastal boundary layer to winds and waves: Analysis of an experiment in Lake Erie. Proceedings, Third Workshop on Great Lakes Coastal Erosion and Sedimentation, National Research Council Canada and Associate Committee for Research on Shoreline Erosion and Sedimentation, 51-54 (1982).

No abstract.

NALEPA, T.F., W.S. GARDNER, and J.M. MALCZYK. Phosphorus release by three kinds of benthic invertebrates: Effects of substrate and water medium. Canadian Journal of Fisheries and Aquatic Sciences 40(6):810-813 (1983).

The effects of a sand substrate (presence vs. absence) and type of water medium (distilled or lake) on phosphorus excretion rates of tubificids, chironomids, and the amphipod Pontoporeia hoyi were determined. In contrast to previous studies on respiration rates, the presence or absence of a substrate did not significantly affect the excretion rates of any of the three taxa. Realistic determinations of P excretion can thus be obtained without a substrate present; this simplifies the approach to such determinations. Excretion rates of tubificids and chironomids were not affected by the type of medium, but the excretion rate of P. hoyi was slightly (but significantly) higher in distilled water than in lake water.

National Research Council, PanelonNiagara River Ice Boom Investigations. The Lake Erie-Niagara River Ice Boom: Operations and Impacts. National Academy Press, Washington, DC, (1983).

No abstract.

PICKETT, R.L. Great Lakes bottom currents. Coastal Oceanography and Climatology News 5:41-43 (1983).

Studies have shown that many contaminants in the Great Lakes attach to particles and settle to the bottom (e.g., Frank et al., 1979). Once these attached contaminants are on the bottom, they are mixed into the sediments by benthic organisms and eventually covered by fresh sediment. Storms, however, upset this process. Fig. 1 shows what happened during a storm over southern Lake Huron on March 26, 1983. Strong winds produced strong currents that extended to the bottom, resuspending sediment and the associated contaminants. This resuspension occurs in all the Great Lakes. The purpose of this note is to examine the frequency and strength of near-bottom currents that cause resuspension in the Great Lakes.

PICKETT, R.L., J.E. CAMPBELL, A.H. CLITES, and R.M. Partridge. Satellite-tracked current drifters in Lake Michigan. Journal of Great Lakes Research 9(1):106-108 (1983). https://www.glerl.noaa.gov/pubs/fulltext/1983/19830002.pdf

Satellite-tracked current drifters are being used to monitor near-surface currents in Lake Michigan. These drifters are now commercially available, and preliminary tests show their satellite-determined positions to be within 0.5 km. The drifters appear to be ideal for monitoring near-surface lake currents and testing hydrodynamic lake models.

PICKETT, R.L., R.M. Partridge, A.H. CLITES, and J.E. CAMPBELL. Great Lakes satellite-tracked current drifters. 1983 Symposium on Buoy Technology, New Orleans, LA, April 27-29, 1983. The Marine Technology Society, 138-143 (1983).

Satellite-tracked current drifters proved to be a valuable tool for monitoring near-surface currents in the Great Lakes. Position accuracy tests conducted at four land locations around the lakes showed that errors were less than 1 km, and did not vary with latitude, longitude, transmitter age, or battery condition. Wind drift was estimated to be 0.7% and was verified with dye and oil spills. Various drogue configurations did not alter drifter tracks. Drifter recovery was aided by the design of a 10-km range, portable radio direction-finder. A replaceable battery module and armor plating enabled reuse of the drifters. Lake Michigan tracks to date show strong northward currents in the east, southward currents in the west, and meandering currents near the middle of the lake.

QUIGLEY, M.A. Freshwater macroinvertebrates. Journal of Water Pollution Control Federation 55:833-840 (1983).

No abstract.

QUINN, F.H., and R.N. KELLEY. Great Lakes monthly hydrologic data. NOAA Data Report ERL GLERL-26, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB84-114545) 79 pp. (1983).

Accurate values of monthly hydrologic data are required for simulation, forecasting, and water resource studies of the Great Lakes and their basins. This report summarizes the monthly hydrologic data currently used by the Great Lakes Environmental Research Laboratory in their hydrologic and water resource studies of the Great Lakes. The data consist of precipitation, runoff, evaporation, connecting channel flows, diversions, beginning-of-month lake levels, and rates of change in storage.

QUINN, F.H., and D.C. NORTON. Great Lakes precipitation by months, 1900-80. NOAA Data Report ERL GLERL-20, Great Lakes Environmental Research Laboratory, Ann Arbor, MI (PB83-175737) 24 pp. (1982).

Accurate values of monthly precipitation are required for simulation, forecasting, and water resource studies of the Great Lakes and their basins. There are often significant errors in the present method of computing these values because of technique problems accentuated when data from preselected stations are missing. Therefore, a monthly precipitation climatology was derived by a modified Thiessen approach using a grid-square technique. The resulting data set, which represents a major improvement over that presently in use, is presented here.

**Reckhow, K.H., and S.C. CHAPRA. Engineering Approaches for Lake Management, Vol. 1: Data Analysis and Empirical Modeling. Butterworth Publishers, Boston, MA, 340 pp. (1983).

No abstract.

Rice, C.P., B.J. EADIE, and K.M. Erstfeld. Enrichment of PCBs in Lake Michigan surface films. Journal of Great Lakes Research 8(2):265-270 (1982). https://www.glerl.noaa.gov/pubs/fulltext/1982/19820003.pdf

The processes of exchange of PCBs and related organics at the surface of large bodies of water such as the Great Lakes are unavoidably linked to the unique properties of surface films. We have observed a 3 to 8 fold enrichment of PCB over underlying water in surface film samples taken from Lake Michigan with a Garrett screen sampler. The particle association of the PCBs was highest in the microlayer, averaging 50% compared to 20% on particles in the subsurface water and less than 5% in the air. The PCBs in the samples were identified as Aroclor 1242. The percentage composition of the two Aroclors varied between the air and the water as follows: the air was 30.2% Aroclor 1254 and the water (microlayer plus subsurface) was 57% Aroclor 1254.

ROBBINS, J.A., and B.J. EADIE. Models of contaminants in Great Lakes. Research in Ocean Engineering: University Sources and Resources 4:3 (1982).

No abstract.

ROBERTSON, A. Lakes--Coastal ecology. In Encylopedia of Beaches and Coastal Environments, M.L. Schwartz (ed.). Hutchinson-Ross Publishing Company, Stroudsburg, PA, 506-508 (1982).

No abstract.

Rumer, R.R., Jr., W.F. Bialas, F.H. QUINN, R.A. ASSEL, and D.W. GASKILL. Niagara River ice boom: Effects on environment. Journal of Technical Topics in Civil Engineering 109:105-116 (1983).

Two approaches are described for examining the possible effect of an ice boom on the air and water temperatures. The first approach utilizes a mathematical model for the ice dissipation process incorporating ice discharge from Lake Erie into the Niagara River. The model provides a useful framework for judging the relative significance of in-lake ice melt and river ice discharge in terms of ice area reduction in the lake. The second approach utilizes statistical testing of air and water temperature data to detect a boom effect. Although the power of the test is limited by the size of the data set, it is concluded that the ice boom has no significant effect on the air and water temperature.

SCAVIA, D. Use and limits of ecosystem models: A Great Lakes perspective. Proceedings, Workshop on Marine Ecosystem Modeling, K.W. Turgeon (ed.), National Oceanic and Atmospheric Administration, Washington, DC, 57-88 (1983).

No abstract.

SCAVIA, D., and W.S. GARDNER. Kinetics of nitrogen and phosphorus release in varying food supplies by Daphnia magna. Hydrobiologia 96:105-111 (1982).

Rates of nitrogen and phosphorus release from individual Daphnia magna were determined by measuring ammonia and soluble reactive phosphorus in successive 10-min incubations in small (0.05 ml) vessels after the animals were removed from their food. Release rates of both nutrients were generally highest initially and decreased with time after removal. The ratio of nitrogen to phosphorus released increased with time after animals were removed from an artificial detritus/bacterial food; ratios were lower and changed with time less for animals fed algae. These data suggest errors bay be introduced by assumptions of constant stoichiometry for nutrient release in varying environments.

SCHWAB, D.J., and K. Hutter. Barotropic and baroclinic eigenmodes of Lake of Zurich and Lake of Lugano. Versuchsanstalt fur Wasserbau, Hydrologie und Glaziologie Internal Report I 64 (1982).

Results of numerical calculations for the barotropic and baroclinic eigenmodes (seiches) of Lake of Zurich and Lake of Lugano are presented. Basic equations for the barotropic modes are the (non-rotating) shallow water equations. Seiche periods and mode structures (surface elevation and current vectors) are calculated by approximating Lake of Zurich and Lake of Lugano with a set of square grids. The baroclinic modes are calculated by using a two-layered-variable depth model and a two-layered variable equivalent depth model. Seiche periods and mode structures are again calculated by dividing the lake domain into a system of grids, but the lake domains are now defined by the intersection of the thermocline depth with the bottom topography. The results are not further interpreted.

Spacie, A., P.F. LANDRUM, and G.J. Leversee. Uptake, depuration, and biotransformation of anthracene and benzo(a)pyrene in bluegill sunfish. Ecotoxicology and Environmental Safety 7:330-341 (1983).

Bluegills (Lepomis macrochirus Raf.) were exposed to [14C]anthracene or [14C]benzo[a]pyrene (B[a]P) in water. Rates of uptake and biotransformation within the fish were followed by 14C counting and thin-layer and liquid chromatography. The initial uptake-rate coefficient for anthracene (KU=36 hr-1) was found to be independent of exposure cocentration. The presence of dissolved humics did not affect anthracene uptake but did reduce the B[a]P uptake rate significnatly. Biotransformation of the anthracene was constant at 0.22 nmol/g/hr, with approximately 92% of the residue unmetabolized at 4 hr. Uptake of B[a]P was linear (KU = 49 hr-1) although biotransformation increased from 0.044 to 0.088 nmol/g/hr between 1 and 2 hr of exposure. Only 11% of the B[a]P 14C activity at 4 hr represented the parent compound. Although 6% of the anthracene was found in liver and gall bladder, 25% of the B[a]P was distributed in the two organs. Depuration rates were first order and yielded half-lives of 17 hr for anthracene and 67 hr for B[a]P. The estimated bioconcentration factors (BCF) for anthracene and B[a]P in whole fish (KU/KP) were 900 and 4900, respectively, for total 14C activity, but only 675 and 490 for parent material. These BCFs were considerable lower than those predicted from the octanol-water partition coefficients because of biotransformation.

Steinhart, C.E., L.J. Schierow, and W.C. SONZOGNI. An environmental quality index for the Great Lakes. Water Resources Bulletin 18(6):1025-1031 (1982).

To facilitate communication on the environmental quality of lakes, particularly among policy makers and the general public, a new index for summarizing technical information is presented . The index is designed for the nearshore waters of the North American Great Lakes, but the concept is applicable to other temperate lakes with relatively good water quality. The index is based on none physical (P), chemical (C), biological (B), and toxic substance (T) variables. Raw data are converted to subindex values by mathematically defined functions based on national or internal objectives. Subindex values are multiplied by weighting factors and added to yield a final score ranging from 0, worst quality, to 100, highest quality. Letters with subscripts following the index score indicate the types and numbers of variables whose values are equal to or worse than the objectives - e.g., 70 C1P1 indicates that one chemical and one physical variable exceeded the objective. For 18 nearshore locations in the Great Lakes, index scores ranged from 98 at two stations in Lake Superior to 30 C2P1B2T3 off Point Mouillee in Lake Erie. If properly utilized, the index should be a useful tool to help managers evaluate the response of the Great Lakes to the multibillion dollar cleanup efforts conducted during the 1970's.

TARAPCHAK, S.J. Soluble reactive phosphorus measurements in lake water: Evidence of molybdate-enhanced hydrolysis. Journal of Environmental Quality 12(1):105-108 (1983).

Measurements of ortho-phosphate (PO4-P) concentrations in natural waters obtained by molybdenum blue methodology are based on the assumption that molybdate (Mo) complexes only the PO4-P (in the absence of arsenate and silicate interference) to form the reductant-sensitive complex 12-molybdophophoric acid (12-MPA). The hypothesis that Mo causes or accelerates PO4-P release from bound sources or forms reductant-sensitive complexes with organic compounds (organic-PO4-MO) was tested by exposing filtered Lake Michigan water to acid, acid plus Mo simultaneously, and Mo, respectively, before measuring PO4-P by the Chamberlain-Shapiro extraction method. Not only does Mo accelerate hydrolysis in the presence of acid, but it also either causes hydrolysis in the presence of acid, but it also either causes hydrolysis or forms organic-PO4-Mo before samples are acidified. Although the relative amounts of 12-MPA and organic-PO4-MO formed in lake water cannot be assessed quantitatively, a major fraction of the blue color formed during routine analysis by this or by similar methods apparently is not the product of acid hydrolysis, but may also be due to reduction of organic-PO4-Mo. Future developments in molybdnum blue methodology aimed at minimizing hydrolysis must consider the sequence of reagent additions and Mo, as well as acid, contact times.

TARAPCHAK, S.J., S.M. BIGELOW, and C. RUBITSCHUN. Soluble reactive phosphorus measurements in Lake Michigan: Filtration artifacts. Journal of Great Lakes Research 8(3):550-557 (1982). https://www.glerl.noaa.gov/pubs/fulltext/1982/19820004.pdf

Tests based on variations in vacuum pressure, volume of sample filtered, and filter pore size demonstrate that filtration artifacts seriously bias estimates of soluble reactive phosphorus (SRP) in water from Lake Michigan. Variations in vacuum pressure and filter pore size can bias estimates by approximately 100%, and variations in sample volume alone can produce nearly a twofold difference in SRP values. These biases are caused by differences in the relative amounts of molybdate-reactive PO4-P released from particulate material and by the retention of P-containing "particles" by filters as pores clog with debris, and are partially responsible for method-specific differences in SRP estimation. Vacuum pressures </= 300 mm Hg, small sample volumes (</= 100 mL), and filters of the same type and pore size should be used for routine SRP estimation.

TARAPCHAK, S.J., R.L. CHAMBERS, and S.M. BIGELOW. Soluble reactive phosphorus measurements in Lake Michigan: Causes of method-specific differences. Journal of Great Lakes Research 8(4):700-710 (1982). https://www.glerl.noaa.gov/pubs/fulltext/1982/19820005.pdf

Method-specific differences in soluble reactive phosphorus (SRP) determinations are thought to be caused in part by differences in acid strengths and exposure times. This premise was tested by comparing SRP concentrations measured in water from Lake Michigan and the Grand River by three methods differing in acid strength, exposure time, and molybdate concentration. Although longer exposure times often result in higher SRP values, more PO4-P can be released from bound sources in lake water with 0.16 N HCl than with 0.4 HCl. Method-specific differences in SRP values, therefore, rarely are proportional to differences in acid strength or exposure time and will vary with changes in the chemical composition of the SRP pool.

TARAPCHAK, S.J., D.R. SLAVENS, M.A. QUIGLEY, and J.S. TARAPCHAK. Silicon contamination in diatom nutrient enrichment experiments. Canadian Journal of Fisheries and Aquatic Sciences 40(5):657-664 (1983).

Large amounts of biologically available silicon (Si) were released into solution from the walls of Pyrex glass reagent bottles and Erlenmeyer flasks during nutrient bioassay experiments using Lake Michigan water. Photosynthetic rates in short-term (4-7 h) incubations and diatom growth rates and maximum yields in long-term incubations (~7 d) were affected by these extraneous Si supplies. The results of conventional nutrient enrichment bioassays performed in low-Si environments can be seriously biased unless Si contamination from glass incubation containers is avoided.

VANDERPLOEG, H.A., and D. SCAVIA. Misconceptions about estimating prey preference. Canadian Journal of Fisheries and Aquatic Sciences 40:248-250 (1983).

No abstract.

Ordering GLERL Publications:

To order a copy of GLERL publications not available for downloading at this site, please contact:

Nicole Rice
Publications Office
NOAA Great Lakes Environmental Research Laboratory
4840 S. State Rd.
Ann Arbor, MI 48108-9719 USA
734-741-2247
734-741-2055 (FAX)