Sie sind auf Seite 1von 187

NON-TARGET IMPACTS OF THE HERBICIDE GLYPHOSATE

A COMPENDIUM OF REFERENCES AND ABSTRACTS

4TH EDITION

INFORMATION REPORT APPLIED MAMMAL RESEARCH INSTITUTE

RATIONALE AND BACKGROUND


The original concept of a compendium of references and abstracts outlining the non-target impacts of the herbicide glyphosate arose from the apparent incomplete and scattered sources of information on this subject. A common complaint from both lay and professional people is: What research has been done on non-target impacts of glyphosate and how do we access this information? In fact, from the computerized literature search which was conducted to identify studies of non-target impacts of glyphosate, the information in this fourth edition of the compendium was extracted from several thousand references covering environmental impacts, toxicology, efficacy, and human health. Thus, there is considerable literature base for glyphosate and this compendium evolved as a means of providing, in as complete a manner as possible, a collection of titles and abstracts of articles reporting on the non-target impacts of this herbicide. As compilers of this document, we have conducted research on the non-target effects of glyphosate over the past 18 years. This work has focused primarily on small mammal populations in forestry and agriculture. Additional work was conducted on black-tailed deer, fish, daphnids, and diatoms (algae) as part of a major field study. To date, with coworkers, there are 18 journal publications outlining our work on the non-target impacts of glyphosate. Much of our earlier work on mammals is summarized in the chapter Effects of Glyphosate on Selected Species of Wildlife: from the book The Herbicide Glyphosate published in 1985.

Druscilla S. Sullivan, M.Sc. Research Associate Thomas P. Sullivan, Ph.D. Director and Research Scientist Applied Mammal Research Institute 11010 Mitchell Avenue, R.R. #3, Summerland, B.C., Canada V0H 1Z0 February 1997

ii

ORIGIN AND USE OF COMPENDIUM


This compendium is designed as an Information Report to provide an objective assessment of the effects of glyphosate on non-target species and ecosystems. The majority of references (scientific journals and proceedings of symposia) up to and including 1993 were extracted from a computerized literature search of the following sources: AGRICOLA (Agriculture 1970 - 1990) CHEMICAL ABSTRACTS (American Chemical Society 1973 - 1993) AQUAREF (Canadian Water Resources References 1970 - 1992) ASFA (Aquatic Sciences and Fisheries Abstracts 1978 - 1993) BIOSIS (Biological Abstracts 1969 - 1993) CAB (Commonwealth Agriculture Bureau Abstracts 1972 - 1993) NTIS (National (U.S.) Technical Information Service 1964 - 1990) The literature search via DC-ROM for this edition was from: BIOSIS (Biological Abstracts 1993 - 1996) LIFE SCIENCES COLLECTION (1993 - 1996) MICROLOG (1993 - 1996) MEDLINE (1966 - 1996) NIOSH (U.S. National Institute for Occupational Safety and Health 1966 - 1996) UKIH (United Kingdom Industrial Health 1966 - 1996)

Titles and abstracts have been reproduced exactly as they appear in the original article or as abstracted by the source system. In the case of journal articles without a formal abstract, a summary of the study has been abstracted by the compilers and this is clearly indicated by an asterisk (*). The compendium is composed of ten sections: Aquatic Intertebrates and Algae, Biodiversity Conservation and Habitat Restoration/Alteration, Birds, Fish, Human Health, Mammals, Microflora and Fungi, Plant and Soil Residues, Terrestrial Invertebrates, and Water Quality. All titles and abstracts of pertinent references with author(s) and publication outlet are listed alphabetically in each section. Some references do not have abstracts and represent those publications which could not be obtained by the compilers. References from outside North American are identified by country, when known, to assist the reader. References which report on species representing more than one section (e.g. fish and aquatic invertebrates) will appear in each section. We recommend that users do not cite information from the abstracts, but rather they refer to the original source, if at all possible.

iii

TABLE OF CONTENTS
I. II. AQUATIC INVERTEBRATES AND ALGAE ......................................................1 BIODIVERSITY, CONSERVATION, AND HABITAT RESTORATION/ALTERATION .......................................................................12 BIRDS ..............................................................................................................37 FISH .................................................................................................................46 HUMAN HEALTH.............................................................................................57 MAMMALS .......................................................................................................72 MICROFLORA AND FUNGI.............................................................................96 PLANT AND SOIL RESIDUES ......................................................................127 TERRESTRIAL INVERTEBRATES ...............................................................160 WATER QUALITY..........................................................................................171

III. IV. V. VI. VII. VIII. IX. X.

iv

Aquatic Invertebrates and Algae

Aquatic Invertebrates and Algae


1. Anton, F. A., M. Ariz, M. Alia, W. Sloof, and H. deKruijf (editors). 1993. Ecotoxic effects of four herbicides (glyphosate, alachlor, chlortoluron and isoproturon) on the algae Chlorella pyrenoidosa Chick. Proceedings of the Second European Conference on Ecotoxicology. Science Total Environment , pp. 845-51 . Herbicides are chemicals used in crop protection to control weeds that could affect freshwater microalgae of natural ecosystems. In this work the ecotoxic effects of several herbicides on the Chlorophycea Chlorella pyrenoidosa Chick. (Chlorella emersonii var. globosa) have been studied using an ecotoxicological bioassay. All herbicides (glyphosate, chlortoluron with terbutryne, isoproturon and alachlor) were commercial and technical products. The EC sub(50) and NOEC values (96-h) of herbicides were calculated to reveal the toxic effects on microalgae caused by them. Isoproturon (50%) and chlortoluron (43%) with terbutryne (7%) were the more toxic herbicides to Chlorella pyrenoidosa. The lower ecotoxic effects that inhibit the algal populations had occurred when algae were exposed to technical and commercial glyphosate. Commercial alachlor and technical and commercial chlortoluron had values between the most and least ecotoxic herbicides that were tested. (Spain) . Austin, A. P., G. E. Harris, and W. P. Lucey. 1991. Impact of an organophosphate herbicide (glyphosateR) on periphyton communities developed in experimental streams. Bulletin of Environmental Contamination and Toxicology 47: 29-35. (*) The effects of low-level concentrations of glyphosate (1-300 g/L) and nutrient enhancement on algal biomass in field-based stream troughs were studied. A 2-week period preceded measurable biomass in all troughs. Although the magnitude of biomass accrual differed between control troughs and those receiving glyphosate, all troughs exhibited biomass increase:decrease cycles. Except for the initial cycle, biomass in the control troughs was less than in those receiving glyphosate. The growth in control+nutrients recovered slowly prior to glyphosate addition following the first cycle, unlike nutrient only troughs. At all three glyphosate concentrations, post-glyphosate growth decreases were followed by biomass accrual values larger than in either control. An assessment of algal species composition indicates all numerically dominant species were present in all troughs throughout the experiment; some patchiness was observed over time, between troughs and treatments. However, with the exception of N. acicularis, no species were lost following the addition of glyphosate. Numerically, diatoms were most abundant, with Achnanthes minutissima increasing to dominate all troughs, although representing only a minor proportion of biovolume. Glyphosate did not appear to inhibit biomass accrual nor was it lethal to any of the pre-treatment dominant species. In troughs receiving the biocide, enhanced growth above control values suggested the phosphate constituent of this herbicide might be acting as a nutrient. The addition of glyphosate to a periphyton community appears to have little effect on subsequent successional patterns. If the primary producer component of periphyton biofilm can use glyphosate or the surfactant as an alternate source of phosphorus, then, irrespective of other toxicological considerations, glyphosate-induced eutrophication of coastal oligotrophic waterways could indirectly affect salmonid habitat and other aquatic resource management concerns. The concentrations used in this study are within the range known to occur in lotic habitats following precipitation events or as a result of direct application. Buhl, K. J., and N. L. Faerber. 1989. Acute toxicity of selected herbicides and surfactants to larvae of the midge Chironomus riparins. Archives of Environmental Contamination and Toxicology 18: 530-536. The acute toxicities of Eradicane (EPTC), Fargo (tri-allate), Lasso (alachlor), ME4 Brominal (bromoxynil), Ramrod (propachlor), Rodeo (glyphosate), Sencor (metribuzin), and Sutan (+) (butylate) and 2 surfactants (Activator N.F. and Ortho X-77) to early fourth instar larvae of C. riparins were determined under static conditions. In addition, technical grade alachlor, tri-allate, metribuzin and propachlor were tested for comparison with the formal dated products. The relative toxicity of the commercial formulations varied considerably. EC50 values ranged from 1.23 mg/l for Fargo to 5600 mg for Rodeo. Fargo, ME4 Brominal and Ramrod were moderately toxic to midge larvae. Lasso,

2.

3.

Aquatic Invertebrates and Algae Sutan (+) and Eradicane were slightly toxic; and Sencor and Rodeo were practically non-toxic. The 48h EC50 values of the two surfactants were nearly identical and were moderately toxic to midges. For two of the herbicides in which the technical grade material was tested, the inert ingredient in the formulations had a significant effect on the toxicity of the active ingredient. Fargo was twice as toxic as technical grade tri-allate, whereas Sencor was considerably less toxic than technical grade metribuzin. A comparison of the slope function values indicated that the toxic action of all the compounds occurred within a relatively narrow range. In general the order of toxicity to C. riparins was similar to those for other freshwater invertebrates and fish. A comparison between estimated max. herbicide concentration in runoff and results of acute tests indicated that Ramrod, ME4 Brominal and Lasso pose the greatest direct risk to midge larvae during a storm event. 4. 5. Buikema, A. L., E. F. Benfield, and B. R. Niederlehner. 1981. Effects of pollution on freshwater invertebrates. Journal of the Water Pollution Control Federation 53: 1007-15. Christy, S. L., E. P. Karlander, and J. V. Parochetti. 1981. Effects of glyphosate on the growth rate of Chlorella. Weed Science 29: 5-7. Growth rates, in terms of cell doublings per day, were determined for cells of C. sorokiniana inoculated into media containing from 5.91 x l0 sup-sup 6M to 591 x 10 sup-sup 6M glyphosate. Comparisons with a control, which supported 10.4 doublings/day, showed growth only slightly reduced at 5.91 x 10 sup-sup 6M and 11.8 x 10 sup-sup 6M, with averages of 9.7 and 9.5 doublings/day; reduced by more than half at 17.7 x 10 sup-sup 6M with an average of 4.4 doublings/day; and prevented at all higher concentrations. In addition to completely preventing growth at 23.7 x 10 sup-sup 6M and above, glyphosate seemed to cause a deterioration of the inoculum cells. Filtering the media containing 5.91 x 10 sup-sup 6M 11.8 x 10 sup-sup 6M, and 17.7 x 10 sup-sup 6M through kaolinite prior to introduction of the alga reduced the inhibition previously noted at 17.7 x 10 sup-sup 6M and resulted in growth rates from 9.1 to 9.5 doublings/day. Environment Canada. 1990. Pesticide research and monitoring annual report 1988-1989. Environment Canada , ISBN 0-662-18461-0; Publication no: DSS cat no En 40-11-13-1989E. See Water Quality Section. Ernst, W. R., R. Morash, W. Freedman, and K. Fletcher. 1987. "Canada Surveillance Report ." Measurement of the environmental effects associated with forestry use of Roundup. EP-5-AR87-8. See Plant and Soil Residues Section. Faust, M., R. Altenburger, W. Boedeker, and L. H. Grimme. 1994. Algal toxicity of binary combinations of pesticides. Bulletin of Environmental Contamination and Toxicology 53 , no. 1: 134-41 . (Germany) . Folmar, L. C. 1978. Avoidance chamber responses of mayfly nymphs exposed to eight herbicides. Bulletin of Environmental Contamination and Toxicology 19: 312-18. In a Y-shaped avoidance maze, mayfly (Ephemerella walkeri) nymphs avoided the highest test concentrations of CuS04 (0.001-0.1 mg CU+2/L, diquat (0.01-1 mg/L), and Roundup (0.1-10 mg/L), but did not avoid 2,4-D DMA (2,4-D dimethylamine salt) (1-100 mg/L), acrolein (0.001-0.1 mg/L), dalapon (0.1-10 mg/L), xylene (0.1-10 mg/L), and Aquathol K (0.1-10 mg/L). The nymphs displayed a marked attraction to 1 mg dalapon/L. The sensitivity of mayfly nymphs to these herbicide pollutants is compared with the sensitivity of rainbow trout fry, and the results are discussed with regard to the impact of herbicide pollution on sport fisheries. Folmar, L. C., H. O. Sanders, and A. M. Julin. 1979. Toxicity of the herbicide glyphosate and several of its formulations to fish and aquatic invertebrates. Archives of Environmental Toxicology and Contamination 8: 269-78. See Fish Section.

6.

7.

8.

9.

10.

Aquatic Invertebrates and Algae 11. Gardner, S. C., and C. E. Grue. 1996. Effects of Rodeo registered and Garlon registered 3A on nontarget wetland species in central Washington. Environmental Toxicology and Chemistry 15, no. 4: 441-51 . See Fish Section. Goldsborough, L. G., and D. J. Brown. 1987. Effects of aerial spraying of forestry herbicides on aquatic ecosystems. Part III. Bioassay of the effect of glyphosate on carbon fixation by intact periphyton communities. Manitoba Environment and Workplace Safety and Health. Water Standards and Studies Report 873. As part of a threeyear study examining potential effects of herbicide use in forest management on adjacent aquatic ecosystems, artificial substrata (acrylic rods) were positioned in shallow water within six experimental study ponds near Pine Falls (96o 10'W, 50o 37'N) during late summer. The substrata were permitted to colonize with periphytic algae for a period of 41 to 50 days, after which samples were collected and exposed in the laboratory to concentrations of glyphosate (added as the commercial formulation Roundup) ranging from 0 to 1800 mg/L. Carbon fixation rates during the first four hours after herbicide treatment were measured, and the change in rate relative to untreated controls was used to assess the shortterm effects of glyphosate on algal photosynthesis. Effects of glyphosate treatment on carbon fixation were not detected in samples from any pond at concentrations less than about 0.89 mg/L. Carbon fixation occurred at the highest tested concentration. Estimated EC50 values for Pine, Hike and Birch Ponds ranged from 35.4 to 69.7 mg/L glyphosate. Based on the estimated glyphosate concentration which could result from a hypothetical application of 2.5 L/ha Roundup to the study ponds (0.06-0.10 mg/L), no immediate effects of herbicide treatment on periphytic algal photosynthesis are predicted. Hartman, W. A., and D. B. Martin. 1984. Effect of suspended bentonite clay on the acute toxicity of glyphosate to Daphnia pulex and Lemna minor. Bulletin of Environmental Contamination and Toxicology 33: 355-61. . 1985. Effects of four agricultural pesticides on Daphnia pulex, Lemna minor, and Potamogeton pectinatus. Bulletin of Environmental Contamination and Toxicology 35: 646-51. The toxicity of glyphosate, atrazine, alachlor and carbofuran to D. pulex, L. minor and P. pectinatus was studied at 0.1-10.0 mg/L and the effect of suspended sediment on the toxicities of these pesticides was also examined. Carbofuran (10 mg/L) had no toxic effects on L. minor with or without suspended sediment. Atrazine and alachlor were toxic to Lemna, and the presence of suspended sediment was relatively unimportant in reducing or increasing the toxicity. The above chemicals had no effect on sprouting of P. pectinatus. Henry, C. J. 1992. "Effects of Rodeo herbicide on aquatic invertebrates and fathead minnows. " M.Sc. Thesis, South Dakota State University. 63 p. (*) I evaluated the effects of Rodeo on the survival of six species of invertebrates and fathead minnows by placing them in enclosures in eight North Dakota wetlands that were aerially treated with Rodeo at 5.8 L/ha and in four that were untreated. The number of animals alive and dead were then counted up to 21 days post-treatment. In field trials, no significant differences (P>0.05) occurred in mortality rates of invertebrates between treated and reference wetlands. Laboratory static acute toxicity testing was done on the same species to determine EC50's or LC50's of Rodeo, X-77 Spreader, and Chem-Trol both individually and in mixtures. Rodeo alone and the field application mixture of all three chemicals were rated as practically nontoxic (100-1000 mg/L), X-77 as moderately toxic (1-10 mg/L), and Chem-Trol as an insignificant hazard (>1000 mg/L). The field application mixture was significantly more toxic than Rodeo by itself with X-77 being the most toxic component. I found no evidence of synergistic effects among the three chemicals. The most toxic EC50 for Rodeo of all species tested was 485 times the highest concentration of glyphosate (0.600 mg/L) found in wetland water samples. I concluded that the benefits to waterfowl by restoring degraded habitat through the use of Rodeo outweigh the limited chance of invertebrate populations being greatly reduced or eliminated as a result of the treatment. Recommendations are given for future research and for managing cattails in wetland complexes with Rodeo herbicide.

12.

13.

14.

15.

Aquatic Invertebrates and Algae 16. Henry, C. J., K. F. Higgins, and K. J. Buhl. 1994. Acute toxicity and hazard assessment of Rodeo, X-77 Spreader, and Chem-Trol to aquatic invertebrates. Archives of Environmental Contamination and Toxicology 27 , no. 3 : 392-99 . The herbicide Rodeo provides waterfowl managers with an effective chemical tool for creating open water habitats in wetlands if its use does not adversely affect native invertebrate communities. The survival of caged Chironomus spp. (midge), Hyalella azteca (amphipod), Stagnicola elodes (pond snail), and Nephelopsis obscura (leech) was assessed in prairie pothole wetlands treated by air with a tank mixture of Rodeo, the surfactant X-77 Spreader, and the drift retardant Chem-Trol at a rate recommended for controlling cattails. Laboratory studies were then conducted to determine the acute toxicities of Rodeo, X-77 Spreader, and Chem-Trol, individually and in simulated tank mixtures, to the same invertebrates and to Daphnia magna in reconstituted water representative of these wetlands. Hernando, F., M. Royuela, A. Munoz-Rueda, and C. Gonzalez-Murua. 1989. Effect of glyphosate on the greening process and photosynthetic metabolism in Chlorella pyrenoidosa. Journal of Plant Physiology 134: 26-31. Chlorella pyrenoidosa pringsheim (211/8 A) was grown photoautotrophically using the herbicide, glyphosate (n-(phosphonomethyl)-glycine), in concentrations ranging from 0.1 mM to 1 mM. Chlorophyll and carotenoid content, greening process, photosynthetic and respiration rates, and photosynthetic electron transport activities were investigated. Glyphosate decreased cell density and photosynthetic pigment content; 1.0 mM did not allow growth. Bleached cells were used for the greening process. Glyphosate had two different effects upon photosynthetic pigments: inhibition of chlorophyll synthesis and a decrease in carotenoids. Oxygen uptake was not affected, but oxygen evolution was strongly inhibited. The results suggest that glyphosate acts as an electron transport inhibitor, acting on both photosystems, but its effect was greater on PS II than PS I. Hildebrand, L. D., D. S. Sullivan, and T. P. Sullivan. 1980. Effects of Roundup herbicide on populations of Daphnia magna in a forest pond. Bulletin of Environmental Contamination and Toxicology 25: 353-57. (*) D. magna populations were arranged in pens in 4 stations in an experimental pond to determine the effects of Roundup (glyphosate). Each station was sprayed with one of the following: pond water alone, field dose (2.2 kg/ha), 10 x field dose and 100 x field dose of Roundup. Survival of D. magna populations was not significantly altered between control and experimental treatments at 2, 4 and 8 days. This study, which attempted to simulate aerial application of Roundup herbicide, indicated that there were no detectable effects on D. magna. Holck, A. R., and C. L. Meek. 1987. Dose-mortality responses of crawfish and mosquitoes to selected pesticides. Journal of the American Mosquito Control Association 3: 407-11. A study was conducted to determine the toxicities (LC50s) of several pesticides on the commercially important red swamp crawfish, Procambarus clarkii, and 3 mosquito species common in Louisiana (U.S.A.) ricelands- Anopheles quadrimaculatus, Culex salinarius and Psorophora columbiae. Pesticides tested in laboratory bioassays included Bacillus sphaericus, B. thuringiensis var. israelensis, Bendiocarb, glyphosate, isostearyl alcohol, Malathion, Propoxur, Resmethrin synergized with Piperonyl Butoxide (PBO) and Thiobencarb. Isostearyl alcohol was the least toxic compound to crawfish, with a LC50 of > 10,000 ppm, while resmethrin + PBO (1:3 ratio) was the most toxic with a LC50 of 0.00082 ppm. The herbicides Glyphosate and Thiobencarb were the least toxic compounds for the mosquito species tested, while B. t. var israelensis and Resmethrin + PBO were the most toxic. Holtby, L. B., and S. J. Baillie. 1989. Effects of the herbicide "Roundup" (glyphosate) on periphyton in Carnation Creek, British Columbia. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 063, pp. 224-31. Periphyton standing crops and accumulation rates were monitored at several sites in the mainstem and valley-bottom tributaries of Carnation Creek for 18 months after the herbicide ROUNDUP was applied at a rate of 2.2 kg glyphosate/ha to much of the valley floor of the watershed. One 4

17.

18.

19.

20.

Aquatic Invertebrates and Algae sidechannel was oversprayed. There is evidence of direct toxicity of some herbicide component two weeks after the mid-September application. In the year following spraying there was some localized enhancement of production as the periphyton responded to increased levels of phosphorus immediately downstream of the over-sprayed side-channel. 21. . 1989. Litter-fall and detrital decomposition rates in a tributary of Carnation Creek, British Columbia, oversprayed with the herbicide Roundup (glyphosate). Forest Resource Development Agreement, B.C. Proceedings of the Carnation Creek Herbicide Workshop. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 232-49. Litter-fall was measured in the riparian zone of a tributary oversprayed with the herbicide ROUNDUP (glyphosate) for 18 months after application. Defoliation of the riparian zone of the tributary reduced deciduous litter-fall to approximately 6% of the expected total deposition of 300 g.m2.yr-1. Coniferous litter was a minor component of all treatment and control sites, averaging less than 1% of the total. At the three control sites coniferous litter-fall increased by an average of 296% in the year after spraying compared to the year before, but in the treatment area the increase was only 36%. Litter decomposition rates in water were variable in the three summers during which measurements were made. Decomposition rates were greatest in the oversprayed tributary and lowest in the control tributary. Between-site differences in decomposition rates were probably the result of differences in their macro-invertebrate communities. Most of the between-year variability in decomposition rates was accounted for by differences in stream temperatures. The herbicide had no measurable effects on macro-invertebrate abundance or community composition for 22 months after application. Decomposition rates in the over-sprayed tributary increased in the summers after spraying, partly as a result of increased summer temperatures that resulted from defoliation of the riparian zone. Hutber, G. N., L. J. Rogers, and A. J. Smith. 1979. Influence of pesticides on the growth of cyanobacteria. Zeitschrist Fur Allgemeine Mikrobiologie 19: 397-402. Two unicellular and 2 filamentous cyanobacteria (blue-green algae) were exposed under conditions optimal for photoautotropic growth to 11 pesticides. Low concentrations (0.01-5 ppm) of diuron, atrazine and paraquat inhibited growth. With MCPA, MCPP, 2,4-D, milstem, and ethrel, marked inhibitory effects were achieved only at concentrations above 100 ppm. Growth was inhibited by glyphosate, DDT, and thiram at intermediate concentrations. In some cases, the effective concentration of the pesticide varied considerably with the organism tested. (Germany) . Kafarov, R. S., L. A. Bakumenko, and D. N. Martorin. 1985. Effect of herbicides on photosynthetic reactions in plants and algae. Agrokhimiya 11: 99-104. (Soviet Union). Kallqvist, T., M. I. Abdel-Hamid, and D. Berge. 1994. Effects of agricultural pesticides on freshwater plankton communities in enclosures. Norwegian Journal of Agricultural Sciences Supplement 13 : 133-52 . (Norway) . Kreutzweiser, D. P., P. D. Kingsbury, and J. C. Feng. 1989. Drift response of stream invertebrates to aerial application of glyphosate. Bulletin of Environmental Contamination and Toxicology 42: 331-38. A study was conducted to determine the drift response of aquatic invertebrates to glyphosate contamination from an aerial application of Roundup (R) to the Carnation Creek watershed in September 1984. Drifting invertebrates were collected from three sampling sites in the watershed and stream water was collected for analysis of residual glyphosate. The sampling sites were located in the main channel of Carnation Creek, in C Creek (a 1000 m ephemeral tributary) and in 1600 Tributary (an 800 m stream directly oversprayed). Glyphosate was applied as Roundup (R) at a concentration of 2.0 kg ai/ha in a total volume of 252 L/ha over four separate days. The application of glyphosate on or adjacent to small tributaries of Carnation Creek did not result in undue disturbance of stream invertebrates. Drift densities of most aquatic invertebrates did not increase in response to the herbicide applications. None of the post-spray mean drift values for total invertebrate catches were significantly 5

22.

23.

24.

25.

Aquatic Invertebrates and Algae higher than pre-spray mean densities. However, the drift response of two organisms, Gammarus sp. and Paraleptophlebia sp., may suggest a slight and ephemeral herbicide induced disturbance in and downstream of the treatment areas. In tributary 1600, residual glyphosate in integrated water samples did not exceed 162 g/L and was less than 50 g/L within 10 h after the overspray. 26. Luetjen, K., I. Girardet, R. Altenburger, M. Faust, and L. H. Grimme. 1988. The effect of glyphosate and phosphinothricin on single celled green algae. Communications of the Federal Biological Institute for Agriculture and Forestry Berlin-Dahlem, 46th German Plant Protection Convention. No. 245. (Germany). Maule, A., and S. J. L. Wright. 1984. Herbicide effects on the population growth of some green algae and cyanobacteria. Journal of Applied Bacteriology 57: 369-79. Six herbicides were tested for their effects on the population growth of a range of green algae and cyanobacteria by an easily replicated low-volume liquid culture technique using Repli-dishes. Diuron, propanil and atrazine were most inhibitory, chlorpropham was intermediate and MCPA and glyphosate were least inhibitory. Chlorpropham was more inhibitory to green algae than to cyanobacteria. The effects of chlorpropham and 3-chloroaniline, a metabolite, on populations of the cyanobacterium Anacystis nidulans and the alga Chlamydomonas reinhardti were monitored in larger scale batch cultures. Both compounds reduced the growth rate although in some cases there was partial recovery. 3-Chloroaniline was less inhibitory than the parent herbicide chlorpropham. (United Kingdom). McLeay, D. 1988. Development of a bioassay protocol for evaluating the toxic risk to regional fisheries resources posed by forest-use herbicides. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 039. 57 p. See Fish Section. Medina, H. S. G., M. E. Loptata, and M. Bacila. 1994. The response of sea urchin egg embryogenesis towards the effect of some pesticides. Arquivos De Biologia e Tecnologia (Curitiba) 37 , no. 4 : 895-906 . Experiments were carried out on the effect of some pesticides on the embryonic development of fertilized eggs from the sea urchin Lytechinus variegatus Lamarck, var. atlanticus. The eggs ere fertilized in a nursery specially designed for such purpose. Three minutes after the formation of the fertilizing membrane, suitable amounts of Trifluraline, Folidol 600, Roundup, Sulfosate and K-Otrine were added to the fertilized egg and their effect observed by microscopic examination. It has been found that specific alterations on the embryonic forms of the egg development were caused by the different pesticides used. The observation began after the addition of the pesticide to the egg suspension and was carried out for 24 hours up to the stage of pluteus. (Brazil) . Nishiuchi, Y. 1978. Toxicity of formulated pesticides to some fresh water organisms. Suisan Zoshoku 25: 148-50. (Japan). Payne, N. J., J. C. Feng, and P. E. Reynolds. 1990. Off-target deposits and buffer zones required around water for aerial glyphosate applications. Pesticide Science 30: 183-98. Off-target deposit has been quantified from various silvicultural glyphosate application methods and an estimate made of the buffer-zone widths required around water to protect fish and their invertebrate food species from possible toxicological effects. To overcome the difficulty of estimating different buffer widths to meet the various use conditions encountered, a realistic worst-case scenario was chosen for small-drop drift and data were collected accordingly. Three glyphosate application methods were tested, employing a helicopter equipped with a 'Microfoil' boom, a 'Thru Valve Boom' and D8-46 hydraulic nozzles respectively. Airborne glyphosate and off-target glyphosate deposits on ground sheets and foliar surfaces were measured at downwind distances between 50 and 200 m from multiple overlaid crosswind swaths. Over this distance airborne glyphosate decreased by factors of 3130, ground deposits by factors of 4-2200, and foliar deposits by factors of 5-100. In general airborne 6

27.

28.

29.

30.

31.

Aquatic Invertebrates and Algae glyphosate and off-target deposit was highest from the D8-46 application, and lowest from the 'Microfoil' boom application. Using these measurements, mathematical equations were formulated to predict glyphosate deposits on water surfaces downwind of multiple swath applications, and calculations made for 100-ha applications. Large-drop drift was also calculated using a ballistic model. An estimate was then made of buffer widths required around water bodies to prevent toxicological effects from small- and large-drop drifts, using reported glyphosate toxicities to salmon, rainbow trout and various aquatic invertebrates. A buffer width of 25 m around water bodies limits mortality in populations of salmon, rainbow trout and aquatic invertebrates to less than 10% for the application methods employing the 'Microfoil' and 'Thru Valve' Boom. For the third application method a 30-m buffer width is suggested. 32. Payne, N., J. Feng, and P. Reynolds. 1989. Off-target deposit measurements and buffer zones required around water for various aerial applications of glyphosate. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 88-109. See Fish Section. Peterson, H. G., C. Boutin, P. A. Martin, K. E. Freemark, N. J. Ruecker, and M. J. Moody. 1994. Aquatic phyto-toxicity of 23 pesticides applied at expected environmental concentrations. Aquatic Toxicology (Amsterdam) 28 , no. 3-4 : 275-92 . Environment Canada uses an Expected Environmental Concentration (EEC) in evaluating the hazard of pesticides to nontarget aquatic organisms. This concentration is calculated by assuming an overspray of a 15 cm deep waterbody at the label application rate. The EEC of pesticides is then related to the EC50 (concentration causing a 50% reduction in a chosen toxicity endpoint) for a given aquatic test organism. At present, the use of an uncertainty factor is suggested in the literature if only a few species are tested because of important interspecific differences in pesticide sensitivity. The phytotoxicity of the EEC of 23 different pesticides to ten algae (24 h inhibition of 14C uptake) and one vascular plant (7-day growth inhibition) was determined in an effort to examine the question of interspecific sensitivity and its relation to the development of pesticide registration guidelines. Chemicals included five triazine herbicides (atrazine, cyanazine, hexazinone, metribuzin, and simazine), four sulfonylurea herbicides (chlorsulfuron, metsulfuron-methyl, ethametsulfuron-methyl, triasulfuron), two phenoxyalkane herbicides (2,4-D and MCPA), two pyridine herbicides (picloram and triclopyr), a substituted urea, an amine derivative, and an imidazolinone herbicide (tebuthiuron, glyphosate and imazethapyr, respectively), a bipyridylium (diquat), a hydroxybenxonitrile (bromoxynil), an aldehyde (acrolein) and an acetanilide (metolachlor) herbicide, as well as two carbamate insecticides (carbofuran and carbaryl) and a triazole derivative fungicide (propiconazole). Test organisms were selected based on ecological relevance and present use in test protocols. Organisms included green algae (Scenedesmus quadricauda and Selenastrum capricornutum), diatoms (Nitzschia sp. and Cyclotella meneghiana), cyanobacteria (Microcystis aeruginosa, Oscillatoria sp., Pseudoanabaena sp., Anabaena inaequalis and Aphanizomenon flos-aquae) and a floating vascular plant, duckweed (Lemna minor). The five triazine herbicides, acreolein and diquat inhibited the carbon uptake of all algae, diatoms and cyanobacteria by more than 50%. Two other pesticides, carbaryl + tebuthiuron caused more than 50% inhibition in 90% of the algae tested. Nine of the 23 pesticides, five of which were triazine herbicides, were therefore highly phytotoxic to algae. Twelve pesticides inhibited growth of duckweed by more than 50%. Once again, all five of the triazine herbicides were among this group, as well as three sulfonylurea herbicides and acrolein, diquat, metolachlor and tebuthiuron. Duckweed was the most sensitive organism tested, being equally affected by all pesticides causing algal phytotoxicity (with the exception of carbaryl), as well as being acutely affected by sulfonylurea herbicides. Green algae were least sensitive to diquat; diatoms and one cyanobacterium were the only organisms that showed sensitivity to glyphosate. Through testing the phytotoxicity of a variety of agricultural pesticides to a wide range of algal taxa, it is evident that there are considerable differences in sensitivity among species and that the use of an uncertainty factor is necessary to provide an acceptable margin of safety in evaluating the hazard presented by these chemicals to the aquatic environment.

33.

Aquatic Invertebrates and Algae 34. Reinbothe, S., A. Nelles, and B. Parthier. 1991. N-Phosphonomethylglycine Glyphosate tolerance in Euglena-gracilis acquired by either overproduced or resistant 5 enolpyruvylshikimate-3-phosphate synthase. European Journal of Biochemistry 198: 365-74. Photautotrophic cells of Euglena gracilis can be adapted to N-(phosphonomethyl)glycine (glyphosate) by cultivation in media with progressively higher concentrations of the herbicide. Two different mechanisms of tolerance to the herbicide were observed. One is characterized by the overproduction and 40-fold accumulation of the target enzyme, 5-enolpyruvylshikimate-3-phosphate synthase, in cells adapted to 6 mM N-(phosphonomethyl)glycine. The other is connected with a herbicide-insensitive enzyme. No evidence was obtained for the involvement of the putative multifunctional arom protein previously reported to be involved in the biosynthesis of aromatic amino acids in Euglena. Cells adapted to N-(phosphonomethyl)glycine excreted shikimate and shikimate 3phosphate into the medium; the amounts depended on the actual concentration of the herbicide. Twodimensional gel electrophoresis and determination of 5-enolpyruvylshikimate-3-phosphate synthase activity in crude extracts, as well as after separation by non-denaturing gel electrophoresis, revealed that the overproduction of the enzyme in adapted cells correlates with the accumulation of a 59-kDa protein. Overproduction of this 59-kDa protein resulted from a selectively increased level of a mRNA coding for a 64.5-kDa polypeptide which appeared in adapted cells, as shown by cell-free translation in the wheat germ system. In contrast to this quantitative, adaptive type of tolerance, the second mechanism causing tolerance to N-(phosphonomethyl)glycine in the Euglena cell line NR 6/50 was probably related to qualitatively altered 5-enolpyruvylshikimate-3-phosphate synthase, which could not be inhibited by even 2 mM N-(phosphonomethyl)glycine in vitro. In agreement with this observation, the putatively mutated cell line excreted neither shikimate nor shikimate 3-phosphate into the growth medium containing N-(phosphonomethyl)glycine, even if cultivated in the presence of 20 mM or 50 mM N-(phosphonomethyl)glycine. (Switzerland). Reinbothe, S., B. Ortel, and B. Parthier. 1993. Overproduction by gene amplification of the multifunctional arom protein confers glyphosate tolerance to a plastid-free mutant of Euglena gracilis. Molecular and Gen. Genetics 239 , no. 3 : 416-24 . Cells of the plastid-free mutant line of Euglena gracilis var. bacillaris, W sub (10)BSmL, can be adapted to glyphosate (N-(phosphonomethyl)- glycine) by gradually increasing the concentration of the herbicide in the culture medium. The molecular basis of glyphosate tolerance is the selective ca. ten fold overproduction of the multifunctional arom protein catalyzing steps 2-6 in the pre-chorismate pathway. Further evidence for an involvement of the multifunctional arom protein is aromatic amino acid synthesis in the plastid-free W sub(10)BSmL cells was obtained by Northern hybridization with AR01-, aroA-, aroL- and aroE- specific Saccharomyces cerevisiae geneprobes encoding the entire arom protein or parts of the EPSP synthase, shikimate:NADP super(+) oxidoreductase and shikimate kinase domains, respectively. Overproduction in adapted relative to control cells of 5.3 kb transcript that cross-hybridized with all of the different probes could be demonstrated. The elevated content of the arom transcript correlated with a selective amplification of two out of five genomic sequences that hybridized with the S. cerevisiae) AR01 gene probe in Southern blots. One of the amplified genomic fragments is assumed to encode the previously identified monofunctional 59 kDa EPSP synthase, which is thought to be an organellar protein, that accumulates to a certain extent in its enzymatically active precursor form of 64.5 kDa in the plastid free W sub(10)BSmL cells. (Switzerland) . Reynolds, P. E., J. C. Scrivener, L. B. Holtby, and P. D. Kingsbury. 1989. A summary of Carnation Creek herbicide study results. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 063, pp. 322-34. See Fish Section. . 1993. Review and synthesis of Carnation Creek Herbicide research. The Forestry Chronicle 69: 323-30. See Water Quality Section.

35.

36.

37.

Aquatic Invertebrates and Algae 38. Richardson, J. T., R. E. Frans, and R. E. Talbert. 1979. Reactions of Euglena gracilis to fluometuron, MSMA, Metribuzin and Glyphosate. Weed Science 27: 619-24. Investigations were conducted on Euglena gracilis Klebs strain Z to determine the effects of fluometuron [1,1-dimethyl-3-( -trifluoro-m-tolyl)urea], MSMA (monosodium methanearsonate), glyphosate [N-(phosphonomethyl)glycine], and metribuzin [4-amino-6-tert-butyl-3-(methylthio)-astriazin-5(4H)-one] on cell number, chlorophyll content, and photosynthesis. Euglena cell number was reduced by 65% or more after 48 h with fluometuron levels above 4 x 10-5 M. MSMA at 6 x 10-4 M reduced cell number 42% after 144 h exposure. Chlorophyll content was reduced 33 to 80% by metribuzin levels of 2 x 10-6 M or greater, and fluometuron inhibited chlorophyll content by 30% or more from 4 x 10-6 M or greater concentrations. Chlorophyll was reduced 21 to 69% by treatment with glyphosate at 3 x 10-3 M, but MSMA appeared to have little effect on chlorophyll except at the high level of 6 x 10-4 M at 48 h. Photosynthesis was reduced 50% or more with metribuzin levels above 9 x 10-7 M and with fluometuron above 9 x 10-5 M. MSMA reduced photosynthesis by 20% at the 6 x 10-3 M level, and glyphosate slightly reduced photosynthesis at levels below 1.2 x 10-4 M but slightly stimulated it above that level. Chronic effects (Euglena exposed to herbicides 96 h prior to measurement) on photosynthesis indicated a more pronounced reduction from fluometuron than from short-term exposure, little change with glyphosate, but less reduction with metribuzin than from shortterm exposure. Metribuzin caused increased respiration rates of 100 to 200% after 100 min of exposure. Respiration was stimulated 20% by glyphosate and relatively unaffected by the other compounds. Removal of Euglena from metribuzin- and fluometuron-treated media to non-treated media resulted in increased levels of chlorophyll to near that of the control. These results suggest that use of these herbicides is not detrimental to non-target algae if the exposure is not intensive. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. Wageningen, NL: European Weed Res. Soc. pp. 315-21. See Water Quality Section. Scrivener, J. C., and S. Carruthers. 1989. Changes in the invertebrate populations of the main stream and back channels of Carnation Creek, British Columbia, following spraying with the herbicide Roundup (glyphosate). Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 263-72. Populations of macroinvertebrates were monitored at two sites in the main stream and at sites with bare mud and mud with rooted vegetation substrates in four tributary swamps. Three tributary swamps (750trib, 1500trib, 1600trib) and one site in the main stream were influenced by the herbicide ROUNDUP (glyphosate) after aerial application. Any impacts from the herbicide were not easily detectable because macroinvertebrate densities varied with substrate types, with the seasons and with previous hydrological conditions. In the swamps, densities of organisms on the surface of the substrate 2 was related to stream flow with a cubic polynomial (R = 0.45). Density maxima occurred during median flows, while density minima occurred during periods with extremes of flow. During periods of freshet, densities in the treated swamp were half those of the untreated swamp. Servize, J. A., R. W. Gordon, and D. W. Martens. 1987. Acute toxicity of Garlon 4 and Roundup herbicides to Salmon, Daphnia, and Trout. Bulletin of Environmental Contamination and Toxicology 39: 15-22. See Fish Section. Shindo, N., Y. Yamamoto, M. Ohyama, and A. Murakami. 1986. Effect of several herbicides against microalgae. Meiji Daigaku Nogakubu Kenkyu Hokoku 71: 918. Growth inhibitor effects of 6 herbicides on microalgae, e.g. Anabaena cylindrica, Chlamydomonas reinhardti, and Cyclotella sp., were investigated. The growth and photosynthesis of the microalgae were inhibited by Me 3,4-dichlorophenylcarbamate, 1,3-dimethyll(5-trifluoromethyl-1,3,4thiadiazol-2-yl) urea, or oxadiazon at the 10 ppm level, but not by 2,4-D or NaCl03. Glyphosate had little effect on respiration at 30 ppm, but slightly reduced photosynthesis. (Japan).

39.

40.

41.

42.

Aquatic Invertebrates and Algae 43. Simenstad, C. A., J. R. Cordell, L. Tear, L. A. Weitkamp, F. L. Paveglio, K. M. Kilbridge, K. L. Fresh, and C. E. Grue. 1996. Use of Rodeo and X-77 spreader to control smooth cordgrass (Spartina alterniflora) in a southwestern Washington estuary: 2. Effects on benthic microflora and invertebrates. Environmental Toxicology and Chemistry 15 , no. 6 : 969-78 . In August 1992, we conducted an intensive short-term (within 119 d) experiment in southern Willapa Bay, Washington, to evaluate the potential effects on mudflat benthic communities of herbicide control of smooth cordgrass, Spartina alterniflora Loisel. A mixture of glyphosate (Rodeo; 4.7 L/ha) and an associated surfactant, alkylarylpolyoxyethylene (AAPOE, X-77 Spreader; 1 L/ha) was applied aerially to three mudflat sites with invasive S. alterniflora. Sediment structure (grain size), edaphic microalgal biomass (chlorophyll a), and densities of benthic and epibenthic meiofauna and benthic macrofauna were sampled systematically in treated and adjacent control (untreated) plots 1 d before, immediately after, and 1, 14, 28, 119 d after spraying. These mudflat biota showed no definitive differences in population trends that would indicate acute responses to the herbicide and surfactant applications over the 119-d duration of the experiment. Two-way ANOVA tests of differences in slope of linear regressions of mean plot microalgal biomass and invertebrate density of 19 taxa groups or species testing short-term (2 weeks) and long-term (17 weeks) trends in response to the experimental treatment tests indicated no significant (p lt 0.1) treatment and only three site effects. Natural variability in the standing stocks (in the case of benthic microalgae) or densities (invertebrates) of most of the 19 indicator taxa prior to spray application was sufficiently high within and between treatment and control plots and among sites to preclude strong inferential tests of acute effects. Although differences in mudflat habitats (e.g., tidal elevation, sediment structure) inherent in the sites prior to treatment affect the power of our ability to test direct effects, there were no indications of either shortor long-term effects on the mudflat community of aerially applying this concentration of herbicide and surfactant. This study did not explicitly address either sublethal or indirect ecological effects, such as associated with an observed decrease in the exotic eelgrass Zostera japonica, which might appear as a longer-term, more subtle response by the mudflat community. Solberg, K. L., and K. F. Higgins. 1993. Effects of glyphosate herbicide on cattails, invertebrates, and waterfowl in South Dakota wetlands. Wildlife Society Bulletin 21 : 299-307 . See Bird Section. Sullivan, D. S., T. P. Sullivan, and T. Bisalputra. 1981. Effects of Roundup herbicide on diatom populations in the aquatic environment of a coastal forest. Bulletin of Environmental Contamination and Toxicology 26: 91-96. (*) This paper reports the effects of RoundupR herbicide (MON 02l39) on diatom populations in several field experiments at the University of British Columbia Research Forest, Maple Ridge, B.C., Canada. A 20-year old Douglas fir plantation received an aerial application of RoundupR herbicide at the recommended rate of 2.2 kg a.i./ha. Sediment samples and slides available for colonization by algae were used to assess the effects of this herbicide on the phytoplankton of a stream and pond. In R another stream Roundup was manually applied at field, l0x and 100x field doses and colonization slides were used to assess toxicity to phytoplankton. It was concluded that variations in abundance of diatoms observed in the ponds and streams were mainly determined by habitat and seasonal factors rather than the aerial or manual application of herbicide. Swinehart, J. H., and M. A. Cheney. 1987. Interactions of organic pollutants with gills of the bivalve molluscs Anodonta californiensis and Mytilus californianus uptake and effect on membrane fluxes II. Comparative Biochemistry and Physiology C. Comparative Pharmacology and Toxicology 88: 293-300. The uptakes of 2,4,5-T, glyphosate, parathion, paranitrophenol, naphthalene, glycine, and inulin by gills of the bivalve molluscs Anodonta californiensis (freshwater) and Mytilus californianus (marine) show non-polar compounds are taken up to a greater extent than polar compounds except where active transport occurs. The uptake of glycine by M. californianus is reduced by pollutants containing complexing functional groups but not by non-polar compounds. The uptake of parathion alters the polyphosphate-inorganic phosphate balance in M. californianus. The uptakes of pollutants parallel their toxicities toward rats.

44.

45.

46.

10

Aquatic Invertebrates and Algae 47. Thomas, M. W., B. M. Judy, W. R. Lower, G. F. Krause, and W. W. Sutton. 1989. Time-dependent toxicity assessment of herbicide contaminated soil using the green alga Selenastrum capricornutum. In: Plants for Toxicity Assessment. W. Wang, J. W. Gorsuch, W. R. Lower (editors). pp. 235-54. Bioassays with S. capricornutum were performed on filtered eluates from a clay loam soil treated 1 h-10 d beforehand with 6 herbicides (glyphosate, imazapyr, triclopyr, picloram, 2,4-D and hexazinone) at recommended rates. The 96-h EC50 (effective concn) values indicated growth inhibition (relative to control sample) for all treatments when assayed 1 h after herbicide application. Algal EC50 values of +100 (control), +27.3 (glyphosate), -20.4 (imazapyr), -22.4 (triclopyr), -49.4 (picloram + 2,4-D) and -100 (hexazinone) were obtained. Assays conducted 10 d after herbicide application to soil revealed substantially reduced toxicity. A slight reduction in toxicity was noted for triclopyr and picloram + 2,4-D, and no change was observed with hexazinone. When the herbicides were applied to water, the following 96-h EC50 (in microg/ml) were obtained: 5300-5500 (imazapyr), 5000 (picloram + 2,4-D), 5000 (triclopyr), 2600 (glyphosate) and 1.2-2.5 (hexazinone). Tooby, T. E. 1985. Fate and biological consequences of glyphosate in the aquatic environment. In: The Herbicide Glyphosate. E. Grossbard, and D. Atkinson, 206-17. London: Butterworths. See Water Quality Section. Trotter, D. M., M. P. Wong, and R. A. Kent. 1990. Canadian water quality guidelines for glyphosate., Water Quality Branch, Inland Waters Directorate, Environment Canada, Ottawa, Ontario. 27 p. See Water Quality Section. U.S.D.I. 1981. Fisheries and Wildlife Research 1980. U.S. Department of the Interior, Fish and Wildlife Service. 201 p. See Fish Section. Wimberely, D. N., and S. G. Berk. 1991. Microplate analysis of dehydrogenase activity in protozoa a rapid aquatic toxicity assay. General Meeting of the American Society for Microbiology 91:321. Wood, D. E. 1996. "Evaluation of the herbicides glyphosate and fluridone for use in waterfowl management impoundments. " M.Sc. thesis , University of Georgia, Athens. See Birds Section.

48.

49.

50.

51.

52.

11

Biodiversity, Conservation, and Habitat Restoration/Alteration

Biodiversity, Conservation, and Habitat Restoration/Alteration


1. Arshad, M. A., K. S. Gill, and G. R. Coy. Barley, canola, and weed growth with decreasing tillage in a cold, semiarid climate. Agronomy Journal 87, no. 1 : 49-55 . Conventional tillage systems are reported to cause soil degradation, yet appropriate conservation tillage practices have not been developed for cold regions of the northern Canadian Prairies. Effects of conventional, reduced, and zero tillage systems (CT, RT, and ZT) on the growth of dryland spring barley (Hordeum vulgare L.), canola (Brassica campestris L.), and weeds were studied on a clay soil (Natriboralf) near Rycroft in northern Alberta. Each crop-tillage combination was fixed in space from 1989 through 1991. Each season, the CT plots were tilled once in the preceding fall and twice in the spring prior to seeding; the RT plots were tilled once prior to seeding in the spring; the ZT plots received a preseeding glyphosate (N- (phosphonomethyl) glycine) application in the spring. Crop residue in ZT was spread evenly by harrowing in the spring just prior to seeding. The 1991 available NO-3-N, NH-4-N, and P in soil or total plant N and P were unaffected by tillage, except that NO-3-N was lower under ZT canola. No consistent effect of tillage was detected on total soil moisture, except for lower moisture in the 0- to 10-cm depth under CT in dry periods. As the study progressed, there was a trend of increased weed population response to tillage, relatively greater weed density under ZT, and a shift in species composition. Mean barley total dry matter (TDM) yield was 3.37, 3.09, and 2.93 mg/ha and grain yield was 1.59, 1.41, and 1.35 mg/ha under RT, ZT, and CT, respectively. Mean canola TDM yield was 2.92, 2.36, and 2.12 mg/ha and grain yield was 0.84, 0.66, and 0.59 mg/ha, under RT, CT, and ZT, respectively. In most cases, however, tillage effects on mean crop yields were nonsignificant (P ltoreq 0.05). Overall, RT was considered to be agronomically and environmentally desirable, due to somewhat better crop yield than ET or ZT systems and two fewer cultivations than CT. Astiz, S., H. Alvarez, and A. Sladeckova. Indirect effects of glyphosate application in the control of Salvinia auriculata in experimental systems. Trav. Assoc. Int. Limnol. Theor. Appl. 25 , no. 4: 2064-67. (Spain). Becker, M., G. Levy, and Y. Lefevre. 1996. Radial growth of mature pedunculate and sessile oaks in response to drainage, fertilization and weeding on acid pseudogley soils. Annales Des Sciences Forestieres (Paris) 53, no. 2-3 : 585-94 . In northeastern France, forest soils on old alluvial terraces are generally unfavourable, strongly acid and often characterized by superficial temporary water tables. In this case, the ground vegetation is dominated by a dense cover of Carex brazoides on the moderately hydromorphic soils (Carex site) or Molinia caerulea on the strongly hydromorphic soils (Molinia site). Both pedunculate and sessile oaks are present in the Molinia site, and practically only pedunculate oak in the Carex site. The experiment aimed at quantifying the radial growth response of mature oaks to various silvicultural interventions. It included i) ditching in order to drain the soils (in 1974), ii) herbicide application (glyphosate; in 1981), and iii) fertilization (P, K, Ca and Mg in 1982; N in 1982 and 1985). A dendrochronological investigation was performed on 620 adult oaks from 60 to 200 years old, which were subjected to these treatments, alone or in combination. The results refer to basal area increment by comparison with control trees. The effect of drainage depended on the site type, the oak species and the age of the trees. Drainage had practically no effect in the Carex site. In the Molinia site, the effect was positive (+20%) for the young (ltoreq 110 years old) sessile oaks only. It became even depressive (-15%) for the old (gt 110 years old) pedunculate oaks. The effect of weeding differed according to the site type, the age of the trees and the drainage modality. Whatever the drainage modality, the effect was depressive (-13%) for the young trees and nonsignificant for the old ones in the Carex site. There was a positive interaction between weeding and drainage in the Molinia site, in the old trees (+22%) as well as in the young ones (+17%), whereas weeding alone had a negative effect (-5%). The effect of fertilization was strongly beneficial (about +20%) in all cases, without any interaction of site type or drainage. However, the time dynamics of this effect was different according to the age of the trees: i) the mean effect was lower in the young trees (+15%), but it was still high when the trees were cored (1991); ii) it was higher in the old trees (+25%), but tended to vanish about 9 years after fertilizing. For 12

2.

3.

Biodiversity, Conservation, and Habitat Restoration/Alteration analyzing the results related to drainage and weeding, we needed to take into account the competition for the mineral nutrients between trees and weeds, as well as the water table depth in the soil, which depends on the evapotranspiration of the whole vegetation cover including trees and ground layer. (France) . 4. Bell, F. W., R. A. Lautenschlager, R. G. Wagner, D. G. Pitt, J. W. Hawkins, and K. R. Ride. 1997. Motor-manual, mechanical, and herbicide release affect early successional vegetation in northwestern Ontario. Forestry Chronicle 73: (In press) . Cover and height of vegetation before and one growing season after: 1) motor-manual cutting, 2) mechanical brush cutting (Silvana Selective/Ford Versatile), 3) aerial application of Release (a.i. triclopyr) herbicide, 4) aerial application of Vision (a.i. glyphosate) herbicide, and 5) control (no treatment) were quantified. Multivariate analysis permitted the study of vegetation response as a whole, while accounting for correlations that exist among the individual vegetation groups. Univariate analysis was used to study the responses of individual vegetation groups. Although no pre-treatment differences in percent cover were observed (P = 0.128), deciduous tree, shrub, forb, grass, and sedge groups responded differently to the treatments after one growing season (P < 0.018). Post-treatment cover of deciduous tree and shrub groups was lower in herbicide treated plots than in cut plots. Forb, grass and sedge covers varied greatly among treatments. Brush saw and Silvana Selective treatments decreased cover of deciduous trees. Release decreased cover of deciduous trees and shrubs. Vision decreased cover of deciduous trees, shrubs and ferns. Cover of all vegetation groups increased on the untreated control. Among the conifer release treatments examined, Vision reduced woody and herbaceous vegetation most. Blackshaw, R. E., and C. W. Lindwall. 1995. Management systems for conservation fallow on the southern Canadian prairies. Canadian Journal of Soil Science 75, no. 1 : 93-99 . . 1995. Species, herbicide and tillage effects on surface crop residue cover during fallow. Canadian Journal of Soil Science 75, no. 4 : 559-65 . Fallow continues to be a common agronomic practice on the Canadian prairies but it has been associated with increased soil erosion. Risk of fallow erosion can be reduced by maintaining adequate levels of crop residue on the soil surface. Field experiments were conducted at Lethbridge, Alberta from 1991 to 1993 to determine if commonly grown prairie crops differ in their rates of crop residue degradation during fallow and to assess the effect of herbicides and wide-blade tillage on loss of crop residues. The ranking of crop residue losses during fallow was lentil gt canola gt rye gt barley gt wheat gt flax. High N content in residues usually increased the rate of biomass loss. Flax straw, perhaps because of its high lignin content, did not follow this pattern and was the most persistent of all crop residues. Up to three applications of the herbicides, glyphosate, paraquat and 2,4-D, at recommended rates did not alter field degradation of any of these crops. These herbicides maintained greater amounts of anchored and total surface crop residues than wide-blade tillage during both fallow seasons. Results are discussed in terms of crops grown before fallow, weed control during fallow, and maintenance of sufficient surface plant residues to reduce the risk of soil erosion. Blixt, D. C. 1993. "Effects of glyphosate-induced habitat alteration on birds, using wetlands." M.S. Thesis , North Dakota State University. Fargo, ND. 127 pp. See Birds Section. Bogart, J. P., R. A. Lautenschlager, and F. W. Bell. 1995. Effects of alternative vegetation management treatments on amphibians and reptiles in the Fallingsnow Ecosystem. Fallingsnow Ecosystem Workshop. Program and Abstracts. Ministry of Natural Resources, Ontario VMAP. pp. 32-33. (*) The effects of different forest release treatments (brush saws, Silvana Selective, Release, Vision, and control) on amphibians and reptiles were examined as part of the Fallingsnow Ecosystem Project. No treatment related population changes have been observed to date. Because of the large area, low herpetofaunal diversity, and non-random distribution of amphibians and reptiles, it was not possible to estimate population densities in the study sites using mark-recapture methods. More

5. 6.

7.

8.

13

Biodiversity, Conservation, and Habitat Restoration/Alteration realistic monitoring techniques, such as estimations of recruitment, age structure, comparative survivorship, and modified levels of heterozygosity, are therefore presently being tested. 9. Boyd, R. S., J. D. Freeman, J. H. Miller, and M. B. Edwards. 1995. Forest herbicide influences on floristic diversity seven years after broadcast pine release treatments in central Georgia, USA. New Forests 10, no. 1 : 17-37 . Maintenance of biodiversity is becoming a goal of forest management. This study determined effects of broadcast pine release herbicide treatments on plant species richness, diversity, and structural proportions seven years after treatment. Three study blocks were established in central Georgia. Plots 0.6-0.8 ha in size were planted to loblolly pine (Pinus taeda L.) in the winter of 1982-83 and then treated with imazapyr (Arsenal), glyphosate (Roundup), and hexazinone (Velpar L. and Pronone 10G) in 1985. In 1992, overstory and understory (lt 1.5 m height) layers were examined utilizing stem and rootstock counts and basal area of overstory species and cover of understory species. ANOVAs were used to test for significance using a randomized complete block model. We found no effect of treatments on species richness. Diversity, measured separately for overstory and understory layers by Shannon-Wiener and Simpson indices, also was not influenced significantly by treatments. Arsenal significantly decreased Diospyros virginiana L. and increased Rubus argutus Link and legumes. Hexazinone treatments generally decreased Quercus nigra L., and Roundup significantly reduced Vaccinium spp. compared to the Check. We concluded that herbicide release treatments did not decrease overstory or understory plant species richness and diversity seven years post-treatment. Caffrey, J. M. 1993. Plant management as an integrated part of Ireland's aquatic resources. Hydroecological Applications 5, no. 1 : 77-96 . This paper outlines the nature and extent of the more important weed problems in Irish aquatic situations and describes a range of control strategies that may be adopted to rehabilitate these habitats and to enable their exploitation by a diversity of user groups. In certain instances paucity of aquatic vegetation presents problems, including bankside destabilization and in channel siltation. Reed transplantation trials have been conducted in denuded Irish canals in an effort to rehabilitate these habitats and the results from these are presented. Weed control procedures generally embrace four broad categories. These are mechanical, chemical, environmental and biological control. Of these only biological control is not widely adopted in Ireland. The effects that these procedures have on specific weeds were quantitatively investigated at a number of aquatic situations throughout the country. Results from experiments among a diversity of aquatic plant species and involving the use of mechanical weed cutting apparatus show that, unless the plant is cut to a depth at which light penetration is significantly diminished, rapid regrowth will occur and higher standing crops than might otherwise be expected may be recorded. Experimentation with a small number of herbicides has revealed that, when used with discrimination, these provide effective weed control, with little or no detrimental effect to the aquatic ecosystem. Results from trials with dichlobaenil and glyphosate on aquatic and riparian species are presented. The effect that shading has on macrophyte biomass and the effect that rotted barley straw has on algal growths in watercourses is also described. Cole, E. C. 1996. Managing for mature habitat in production forests of western Oregon and Washington. Weed Technology 10 : 422-28 . Standard timber management practices in the Pacific Northwest result in stands which often vary from unmanaged stands in structure and composition. Forest and wildlife managers have identified a deficit of stands in the mature (> 100-year-old) age class that contain certain desirable wildlife habitat features. Techniques are being developed that would increase the likelihood that managed stands can produce these characteristics. The key desirable components in these stands include large (> 75 cm diam breast height) conifer trees, snags, coarse woody debris, and understory structure, including regeneration. Vegetation management techniques can facilitate development of these components within stands. Thinning the overstory, underplanting shade-tolerant species, and creating snags and coarse woody debris can be accomplished within a production forest. Maintaining shade-intolerant species requires a higher level of disturbance and canopy opening than needed for shade-tolerant species. Treatments which remove competition from shrubs and herbaceous plants may be necessary to insure growth and survival of understory regeneration. Injection of different herbicides into low-grade conifers may yield different types of snags in comparison to girdling or

10.

11.

14

Biodiversity, Conservation, and Habitat Restoration/Alteration topping. Although much of the understory may be eliminated during future thinnings and final harvest, some of the structure will remain and could be carried over into the next rotation along with snags and large coarse woody debris. These treatments are expected to enhance mature habitats in present and future cycles with minimum impact on yield. 12. Cole, E. C., W. C. McComb, M. Newton, C. L. Chambers, and J. P. Leeming. 1995. Response of small mammal and amphibian capture rates to clearcutting, burning, and glyphosate application in the Oregon coast range. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R. E. Gaskin, and J. A. Zabkiewicz, (compilers). FRI Bulletin No. 192. pp. 155-57. See Mammals Section. . 1997. Response of amphibians to clearcutting, burning, and glyphosate application in the Oregon coast range. Journal of Wildlife Management (In press) . We sampled amphibians on 3 red alder (Alnus rubra) sites 1 year before and 1 and 2 years after the following treatments were applied to each site: (1) control (uncut), (2) clearcut and broadcast burned, and (3) clearcut, broadcast burned, and then sprayed with the herbicide glyphosate. All sites included uncut riparian buffer strips. For 3 of the 6 species with >20 captures in pitfall traps, we did not detect changes in capture rates after clearcutting. Capture rates of ensatina salamanders (Ensatina eschscholtzii) and Pacific giant salamanders (Dicamptodon tenebrosus) decreased after logging. Capture rates of western redback salamanders (Plethodon vehiculum) increased the first year after logging, probably because the salamanders sheltered in pitfalls, but effects on populations were unclear. Logging did not significantly alter capture rates of rough-skin newts (Taricha granulosa), Dunns salamanders (P. dunni), and red-legged frogs (Rana aurora). Planning the location and timing of clearcuts or other silvicultural practices over a landscape and retaining riparian buffer strips may be necessary to ensure long-term persistence of Pacific giant salamanders. We did not detect any effects of herbicide spraying on capture rates. Capture rates for rough-skin newts and red legged frogs were higher in uncut red alder stands than in Douglas-fir (Pseudotsuga menziesii) stands sampled in other published studies, an indication that, when red alder stands are converted to Douglas-fir, some trees of the former should be left adjacent to streams to provide habitat for these species and other hardwood associates. Cole, E. C., W. C. McComb, M. Newton, J. P. Leeming, and C. L. Chambers. 1997. Response of small mammals to clearcutting, burning, and glyphosate application in the Oregon coast range. Journal of Wildlife Management (submitted) . See Mammals Section. Edwards, M. B., and B. D. Shiver. 1993. Can forest site preparation benefit stand-level diversity? Proceedings of the International Conference on Forest Vegetation and Management - Ecology, Practice, and Policy. D. H. Gjerstad (compiler). Auburn University. pp. 120-123. ... A silvicultural perspective is presented which includes a new forest stand-level index for diversity calculations, Silviculture Diversity Index (SDI). ... The SDI incorporates the herbaceous as well as the woody components and presents a more complete description of stand level diversity [including wildlife foods]. The new diversity index as well as ... traditional indices are calculated for 10 different site preparation treatments [a control, an intensive mechanical treatment, and 8 chemical treatments] in the Georgia Piedmont. Results indicate that only two treatments decreased biodiversity after 5 years. The greatest decrease was found in the control - an indication that silvicultural practices can be beneficial in creating and maintaining diversity in young pine stands. All but one ... treatment actually increased biodiversity as compared to initial conditions. Eggestad, M., E. Enge, O. Hjeljord, and V. Sahlgaard. 1988. Glyphosate application in forestecological aspects. VIII. The effect on Black Grouse (Tetrao tetrix) summer habitat. Scandinavian Journal of Forest Research 3: 129-35. See Birds Section.

13.

14.

15.

16.

15

Biodiversity, Conservation, and Habitat Restoration/Alteration 17. Ferm, A., J. Hytonen, S. Lilja, and P. Jylha. 1994. Effects of weed control on the early growth of Betual pendula seedlings established on an agricultural field. Scandinavian Journal of Forest Research 9, no. 4 : 347-59 . Various herbicides (glyphosate, sethoxydim, pendimethalin, chlorthiamid, dichlobenil, terbuthylazine) as well as particle board mulch and a cover crop (clover, Trifolium repens) were compared during the first two post-planting years as weed control means in a silver birch (Betula pendula) plantation established on agricultural soil in southern Finland. Chlorthiamid, dichlobaenil and terbuthylazin exhibited good weed control for two years and also increased the height growth of the seedlings by 40-50 cm and much greater relative increase in the leaf area and volume as compared to untreated control plots. Weed control had a significant effect on the foliar nutrient concentrations of birch, particularly that of N. As the amount of weed vegetation increased, foliar N, P (second year), K, Cu and B decreased and, respectively, foliar P (first year), Ca and Mg increased. Vegetation control also had a great indirect influence on the state of health of the seedlings. Incidences of vole damage and bark necrosis disease were associated with a high cover-percentage of weeds, particularly of clover which is much favored by voles. Particle board mulch seemed to increase vole damage by providing shelter for the voles. The competition by weeds for nutrients, and probably also for water, was much more important than their competition for light. Use of mulch and a cover crop did not reduce root competition as effectively as did the best herbicides. (Finland) . Flickinger, E. L., and G. W. Pendleton. 1994. Bird use of agricultural fields under reduced and conventional tillage in the Texas Panhandle. Wildlife Society Bulletin 22 : 34-42 . `We conducted bird surveys in reduced-tillage and conventional tillage fields in spring, summer, fall, and winter from 1987 to 1991 in the Texas Panhandle. Eastern meadowlarks, longspurs, and savannah sparrows were more common in reduced-tillage (sorghum and wheat stubble) fields than in conventionally tilled (plowed) fields in at least 1 season. Other species also had patterns suggestive of greater abundance in reduced-tillage fields. Horned larks, which prefer habitat with sparse vegetation, were more abundant in plowed fields in all seasons except summer. Bird diversity was greater in reduced-tillage fields than in conventionally tilled fields in summer. Cover density and height were greater in reduced-tillage fields in all seasons except spring. Cover density and height rather than cover composition (e.g., grain stubble or live plants) seemed to be the important factors affecting bird distribution. Patterns of bird abundance between sorghum and wheat stubble fields also were dependent on cover. Herbicide use was not greater in reduced-tillage fields than in conventionally tilled fields. Reduced-tillage agriculture for sorghum and wheat farming should be encouraged in the southern Great Plains as a means of improving the attractiveness of agricultural land to many bird species. Freedman, B. 1991. Controversy over the use of herbicides in forestry, with particular reference to glyphosate usage. Journal of Environmental Science and Health C8: 277-86. See Mammals Section. Gardner, S. C., and C. E. Grue. 1996. Effects of Rodeo registered and Garlon registered 3A on nontarget wetland species in central Washington. Environmental Toxicology and Chemistry 15, no. 4: 441-51 . See Fish Section. Gordon, A. M., J. A. Simpson, and P. A. Williams. 1995. Six-year response of red oak seedlings planted under a shelterwood in central Ontario. Canadian Journal of Forest Research 25, no. 4 : 603-13 . This study investigated the potential for underplanting 1+0 and 1+1 northern red oak (Quercus rubra L.) in conjunction with a two-cut shelterwood harvest of a low-quality, ridgetop, tolerant hardwood stand in central Ontario. Growth patterns were followed for 6 years; herbicide (glyphosate) application and prescribed fire (two burns, 3 years apart) were used as competition control measures in an experimental design that tested all possible combinations of stock type, competition control, and clipping 2 years after planting. Six years after outplanting the mean height attained across all treatments for 1+0 stock was 76 cm and for 1+1 stock, almost 90 cm. Depending upon treatment, some individuals were almost 3 m tall. In the presence of heavy deer (Odocoileus virginianus Zimmerman) 16

18.

19.

20.

21.

Biodiversity, Conservation, and Habitat Restoration/Alteration browsing, competition control of any type does not appear to be warranted (i.e., control treatment). The general technique for establishment of oak in this manner appears to be logical for these site conditions, but further investigations necessary to address the direct impact of browsing pressure on seedling growth and the potential for similar silvicultural prescriptions on more marginal oak sites. 22. Graves, D. H., and J. M. Ringe. 1993. European black alder survival and growth responses to herbicide treatment on an eastern Kentucky vegetated coal surface mine excess spoil area after three and eight years. International Journal of Surface Mining and Reclamation 7, no. 1 : 37-40 . A research area was established on a surface coal mine in eastern Kentucky in 1982 to test the effects of eight herbicide treatments, mechanical scalping and no treatment on the long term survival and growth of European black alder (Alnus glutinosa). After eight years, three treatments (mechanical scalping and glyphosate applied simultaneously with napropamide prior to planting or by using napropamide as a post planting treatment after initial applications of glyphosate as a pre-plant treatment) were superior to one or more of the other treatments for all eight parameters tested and were indifferent to each other. If the parameter, percent growth change over time, was not considered an important factor in establishing this species, no treatment was significantly different from the control. Such results indicate factors other than biological responses should be considered when selecting a woody vegetation establishment scheme on surface coal mine excess spoil areas with existing herbaceous cover. Grilz, P. L., and J. T. Romo. 1995. Management considerations for controlling smooth brome in fescue prairie. Natural Areas Journal 15, no. 2 : 148-56 . Smooth brome (Bromus inermis Leyss.), an introduced perennial grass, is an aggressive invader of prairie dominated by plains rough fescue (Festuca altaica Trin. subsp. hallii [Vasey] Harms). We (1) compared richness and density of plant species in brome and fescue stands that were unburned or burned in spring or fall; (2) determined the effects of wick application of a 33% glyphosate solution, applied when smooth brome was in the boot stage, on the density of brome and native flora in unburned plots and plots burned in spring or fall, and; (3) determined the composition of the seedbank for its potential contribution to natural revegetation following the control of smooth brome. Stem densities of native species and plains rough fescue were about two- and fivefold greater, respectively, in fescue plots than in brome plots. Species richness was generally slightly greater in fescue than in brome plots. Burning had no significant effect on stem density of smooth brome. At one site, changes in the density of smooth brome were affected by the interacting effects of burn treatments and glyphosate application. Glyphosate eliminated brome in the spring burn plots and reduced densities 76 and 50% (SE plus or minus 6.4) in the fall burn and unburned plots, respectively. At a second site, smooth brome densities were reduced by glyphosate but not by burn treatments. There was, however, a trend for greater reduction of smooth brome with glyphosate application in the spring burn (98%) than in the fall burn (40%) and unburned (56%) plots (SE plus or minus 15.5). Glyphosate reduced the density of native graminoids 91% (SE plus or minus 10.0), but plains rough fescue and native forbs were not affected. Twenty-three species emerged from the seedbank in fescue plots, whereas twenty emerged from soils collected in brome plots. The total number of seedlings emerging from the 2 2 seedbank was similar in brome and fescue plots, averaging 1794/m and 2078/m (SE plus or minus 252), respectively. The proportion of seedlings emerging was lower for native graminoids (33 vs. 41%, SE plus or minus 2) and greater for native forbs (56 vs. 48%, SE plus or minus 3) in brome as compared to fescue plots. Smooth brome seedlings emerged only from soils collected in brome plots, averaging 3/m2. These studies indicate that excellent control of smooth brome can be achieved with spring burning followed by wick application of glyphosate; however, native species were also reduced by glyphosate. Additional glyphosate applications will be required for complete control of smooth brome. The number of seeds of native species in the soil in smooth brome stands approximates that in stands of plains rough fescue, suggesting that there is an adequate seedbank for natural recovery of vegetation following control of smooth brome. Gutierrez, E., F. Arreguin, R. Huerto, and P. Saldana. 1994. Aquatic weed control. International Journal of Water Resource Development 10, no. 3: 291-312. See Plant and Soil Residues Section.

23.

24.

17

Biodiversity, Conservation, and Habitat Restoration/Alteration 25. Gutierrez, L. E. 1993. Effect of glyphosate on different densities of water hyacinth. Journal of Aquatic Plant Management 31 : 255-57 . Water hyacinth (Eichhornia crassipes) is a problem in rivers and reservoirs in Mexico and appropriate control methods need to be evaluated. The phytotoxic effects of the herbicide glyphosate were determined on water hyacinth in the reservoir Endho, State of Hidalgo, Mexico. Duplicate 1 m2 areas initially containing 10, 20, 30, and 40 kg/m2 densities of water hyacinth were randomly established. After a 2 week acclimatization period, two doses of glyphosate (Rodeo at 5 l/ha and 7 l/ha) were applied to the plots and plots were evaluated at 9, 19, 33, 51, 72 and 94 days after treatment. Glyphosate at 5 l/ha (2.38 kg/ha acid) effectively controlled densities of 10 and 20 kg/m2 in 51 days. At higher plant densities (30 and 40 kg/m2), a second application of glyphosate at day 51 of 2 l/ha (0.95 kg/ha) was necessary to achieve plant control. The response of the plants indicated that a second application 20 to 30 days after the first would provide good results in high density areas. (Mexico) . Harrington, T. B., R. G. Wagner, S. R. Radosevich, and J. D. Walstad. 1995. Interspecific competition and herbicide injury influence 10-year response of coastal Douglas-fir and associated vegetation to release treatments. Forest Ecology and Management 76 : 55-67 . Responses of competing vegetation and planted Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco var. menziesii) were studied for 10 years after six herbicide and manual release treatments in the Washington and Oregon Coast Ranges. Research objectives were to quantify regional, long-term responses of vegetation to various levels of competition, light and soil water availability, and intensity versus importance of factors influencing Douglas-fir growth. Three treatments reduced shrub cover relative to the untreated check: triclopyr in year 1, glyphosate in years 1-5, and repeated control (via several herbicide applications) in years 1-10. Reductions in woody cover from glyphosate stimulated increases in herb cover in years 3 and 5, while repeated control reduced herb cover in years 1, 2 and 5. Through year 10, Douglas-fir survival (86-99%) varied little among treatments. Visual symptoms of herbicide injury to Douglas-fir from triclopyr (45% of trees) and glyphosate (17% of trees) were associated with 0.1-0.2 m reductions in first-year height. After adjusting for tree size, Douglas-fir growth in stem basal area 2 years after triclopyr was less than that of the untreated check, suggesting prolonged effects of herbicide injury. Because it sustained low levels of interspecific competition, caused minimal tree injury, and prevented overtopping cover from red alder (Alnus rubra Bong.), repeated control was the only treatment in which Douglas-fir size (9.8 m height and 21 cm basal diameter in year 10) significantly exceeded (P<0.02) that of the untreated check (7.8 m height and 12 cm diameter). Henry, C. J. 1992. "Effects of Rodeo herbicide on aquatic invertebrates and fathead minnows. " M.Sc. Thesis, South Dakota State University. 63 p. See Aquatic Invertebrates and Algae Section. Henry, C. J., K. F. Higgins, and K. J. Buhl. 1994. Acute toxicity and hazard assessment of Rodeo, X-77 Spreader, and Chem-Trol to aquatic invertebrates. Archives of Environmental Contamination and Toxicology 27, no. 3 : 392-99 . See Aquatic Invertebrates and Algae Section. Hjeljord, O., V. Sahlgaard, V. E. Enge, M. Eggestad, and S. Gronvold. 1988. Glyphosate application in forest-ecological aspects. VII. The effect on mountain hare (Lepus timidus) use of a forest plantation. Scandinavian Journal of Forest Research 3: 123-27. See Mammals Section. Horsley, S. B. 1994. Regeneration success and plant species diversity of Allegheny hardwood stands after Roundup application and shelterwood cutting. Northern Journal of Applied Forestry 11, no. 4 : 109-16 . The presence of desirable regeneration and plant species diversity was studied in five Allegheny hardwood stands before and for 7 yr after application of Roundup herbicide and shelterwood cutting to remove interfering understories of hayscented and New York fern, striped maple, and beech and to establish desirable hardwood regeneration. In each of five 20 ac stands, 10 ac were sprayed 18

26.

27.

28.

29.

30.

Biodiversity, Conservation, and Habitat Restoration/Alteration with 1 lb ai (active ingredient)/ac of Roundup in August 1979; the remaining 10 ac were unsprayed. The entire 20 ac stand received a shelterwood seed cut the following winter. On treated plots, fern was reduced from 57% stocking before treatments to 7% 1 yr after treatment; striped maple and beech stocking was reduced by 16% during the same time. By the fourth year after treatment, regeneration stocking of desirable hardwood species had doubled on treated plots but remained unchanged on untreated control plots. Species diversity was measured by species richness, three diversity indices Berger-Parker, Margalef, and Shannon, an index-free diversity ordering. Neither the herbicide treatment nor the shelterwood seed cut had a statistically significant effect on species richness of either woody species or herbaceous species groups. Nor was there a significant effect of treatments on woody species diversity as measured by the three diversity indices. Index-free diversity ordering showed: (1) a trend toward increase in less common species on Roundup-treated areas, and (2) the diversity of herbaceous species groups was higher in all 7 yr after treatment with Roundup. Values of the three diversity indices were higher for herbaceous species on Roundup-treated subplots for 2 to 7 yr after treatment. Much of this increase resulted from germination of seedbank grasses, sedges, and Rubus spp. with the qualification that long-term and current browsing by deer has impoverished the flora of the Allegheny hardwood forest, diminishing potential treatment effects, the study demonstrated that Roundup treatment aided desirable regeneration establishment, did not have a negative impact on woody species or herbaceous species group diversity, and may have increased diversity of herbaceous species. 31. Hoy, J. B., D. L. Dahlsten, and P. J. Shea. 1982. The effects of 2,4-D and an alternative material (Glyphosate) used for brushfield rehabilitation on soil arthropod community and litter decomposition rate., Final Report USAUCB, Research Agreement PSW-81-0026. 34 p. Jackson, N. E. 1996. Chemical control of saltcedar (Tamarix ramosissima). Proceedings of the Saltcedar Management Workshop, Imperial County and UC Davis : University of California Cooperative Extension. pp. 21-26. Chemical weed control may be the optimal method for control and removal of salt cedar during the establishment of native habitat restoration projects. Roundup, Rodeo, Arsenal, Garlon 3A and 4, and Pathfinder II herbicides are the products most often chosen by managers of native habitat restoration projects because these herbicides are efficacious and cost-effective, with favorable environmental and toxicological properties. Herbicides may be used effectively, within a defined management plan. Individual products may be applied as a foliar treatment by hand or aerially, or as a Basal Bark or Cut-Stump treatment, based on label recommendations. Post-spray management includes retreatment of escapes and new growth in subsequent years, while saving desirable plants, protecting animal life, enhancing wildlife habitat and protecting water quality in riparian, estuarine and terrestrial areas. The most desirable weed management in native habitat restoration projects may utilize a combination of chemical, mechanical, biological and competitive methods. Long-term, healthy competition from the desired species, coupled with chemical control of any re-invading exotic plants may be the optimal program. In any given project, the best combination of tools should be selected and molded into a viable weed management program. Karlsson, A. 1996. Initial seedling emergence of hairy birch and silver birch on abandoned fields following different site preparation regimes. New Forests 11, no. 2 : 93-123 . Experiments were carried out at four sites in Sweden to investigate the possibility of establishing hairy birch (Betula pubescens Ehrh.) and silver birch (Betula pendula Roth) on abandoned fields using natural regeneration and (or) direct seeding. The effects of six soil preparation methods (no preparation, ordinary ploughing, rotary cultivation, deep ploughing, inverted ground, removal of topsoil) and five additional treatments (no treatment, herbicide, peat litter, wood-ashes, slaked lime) on seedling emergence percentages (SEP) and vegetation cover percentages (VCP) were studied. SEPvalues were estimated in June (SEP(J)) and October (SEP (O)). The experimental designs used were: 1) split-plot design with whole plots in a randomized complete block design (RCBD); 2) RCBD. The SEP-values observed with no preparation were close to 0% while the SEP-values obtained with mechanical soil preparation methods mostly were much higher (p ltoreq 0.014). Seedbeds with top-soil in the surface, created by ordinary ploughing or rotary cultivation, and seedbeds with mainly bare mineral soil in the surface, created by deep ploughing, inverted ground or removal of top-soil, obtained SEP-values of equal merit on silty soils, reaching SEP (0)-values up to 15%. The latter seedbeds 19

32.

33.

Biodiversity, Conservation, and Habitat Restoration/Alteration obtained the best result on sandy soil, with as high SEP (0)-values as 47% after removal of top-soil. Seedbeds with top-soil in the surface were quickly colonized by ground vegetation, reaching VCPvalues between 70% and 100%. Herbicide spraying with glyphosate and application of peat litter to the seedbed surface promoted seedling emergence. However, herbicide spraying before soil preparation was of little effect when followed by mechanical soil preparation. (Sweden) . 34. Kay, S. H. 1995. Efficacy of wipe-on applications of glyphosate and imazapyr on common reed in aquatic sites. Journal of Aquatic Plant Management 33 : 25-26 . Common reed, Phragmites australis (Cav.) Trin., is a helophytic grass found in all coastal states of the United States and has been reported to occur in nearly all states within the continental United States. Reeds are abundant and often dominate the flora in temperate climates around the fringes of marshes, particularly where they grade into freshwater wetlands. The U.S. Fish and Wildlife Service (USFWS) and other agencies, particularly in the Atlantic coastal and Great Lakes states, consider common reed to be a highly-invasive weed having little value for fish and wildlife. Large infestations of reeds usually are managed with herbicide applications, particularly glyphosate [N(phosphonomethyl) glycine]. Weed control following herbicide application is only temporary, so frequent re-treatments are needed. Herbicide spraying frequently eliminates more desirable non-target vegetation. Alternatives to herbicide spraying for reed control such as cutting, burning, or chopping also have serious impacts on nontarget plants. The objectives of this study were to evaluate the feasibility to use wipe-on herbicide technology to control reed growth in a shallow, freshwater marsh environment as a more environmentally-acceptable alternative to spray applications and to compare this technique with mechanical removal. Kilbride, K. M., F. L. Paveglio, and C. E. Grue. 1995. Control of smooth cordgrass with Rodeo in a southwestern Washington estuary. Wildlife Society Bulletin 23, no. 3 : 520-524 . We investigated efficacy of aerial and ground applications of Rodeo to control smooth cordgrass (Spartina alterniflora) on intertidal mudflats in Willapa Bay, Washington during 1992-1993. The herbicidal formula was applied to 2 1-ha plots using a helicopter with tow-mounted spray boom or by hand-held wand sprayer. Stem densities did not differ from pre- to post-treatment between aerially treated and control clones. Stem densities were reduced on clones where the herbicidal formulation was applied by hand. Stem density was reduced most at plots located highest in the intertidal zone. Efficacy depended on the time between treatment and subsequent tidal inundation of treated plants. Krishka, C. S., J. F. Connor, L. J. McMillan, H. R. Timmermann, and J. G. McNicol. 1990. Vegetation dynamics following an application of glyphosate in boreal mixedwood sites of Ontario. Technical Report. Ontario, Ministry of Natural Resources, Northwestern Ontario Forest Technology Development Unit, Ontario. Changes in vegetative communities were studied in four North Central Ontario clearcuts following an application of glyphosate. Each cutover had a comparative untreated control area. The areas were all artificially regenerated with black spruce (Picea mariana). There was some naturally regenerated jack pine (Pinus banksiana). Nineteen shrub and tree species were recorded and the number of stems by height class were measured. Five species were more prevalent in the unsprayed control areas after treatment. Sixty-eight herbaceous genera were assessed for number of plants and percent cover. After treatment, six genera were found to occur with greater percent cover in the control areas, while three genera had greater percent cover in the sprayed area. There were greater numbers of Aster in the unsprayed control areas one year after treatment. However, Aster occurred in greater numbers in the spray treated areas three years after treatment. No significant differences between treated and control areas were found for percent ground cover of sedges (Cyperaceae), grasses (Graminaceae), moss or Scirpus. No differences were recorded for height growth or root collar diameter of black spruce or jack pine two or three years after treatment. Changes in growing conditions after spraying favoured some vegetation and disfavoured others. Recovery of some vegetation was observed within the three-year period. Three years of post-treatment assessment may not have been long enough to measure growth impact for conifers. Vegetation recovery and crop response should be monitored for a longer period of time to determine the long-term changes in vegetation dynamics and crop response.

35.

36.

20

Biodiversity, Conservation, and Habitat Restoration/Alteration 37. Lambert, F., B. Truax, D. Gagnon, and N. Chevrier. 1994. Growth and N nutrition, monitored by enzyme assays, in a hardwood plantation: Effects of mulching materials and glyphosate application. Forest Ecology and Management 70, no. 1-3 : 231-44 . The effects of mulching materials (black plastic and straw) and herbicide application (glyphosate) on growth and nitrogen nutrition of butternut (Juglans cinerea), white ash (Fraxinus americana) and bur oak (Quercus macrocarpa) were studied in a plantation established in an abandoned field in southern Quebec (Canada) since 1987. To ascertain the efficiency of mulching materials on the performance of tree seedlings, a herbicide (glyphosate, 6 l/ha) was applied in half of the experimental plots in June 1990 and 1991. Soil parameters (temperature, moisture) and soil nitrate and ammonium concentrations were measured during the growing season in 1991. In summer 1991, nitrogen nutrition of the seedlings was monitored using enzyme assays (nitrate reductase activity (NRA), glutamine synthetase activity (GSA)) of the leaves of the three deciduous tree species. Mulching affected soil parameters, with black plastic producing the highest soil temperature (23.4 degrees C) and straw the highest soil moisture (183.6 g/kg) in June. NRA varied in relation to tree species, herbicide application, mulching material and time. GSA was poorly correlated to silvicultural treatments. Butternut showed the highest NRA, especially in herbicide plots irrespective of mulching material. Bur oak NRA showed less variation between herbicide and nonherbicide plots and reacted more to the mulching treatments. White ash NRA only showed a herbicide effect in June. All species reacted positively to the addition of a herbicide around the mulching material, but not to the same degree. Height and diameter increment ratios (with/without glyphosate application) indicate that butternut growth was the most improved by herbicide applications, followed by white ash and bur oak. This indicates that the effectiveness of mulching alone decreases in the following order: bur oak gt white ash gt butternut. The utilization of mulching material in abandoned fields as an alternative to herbicide application is closely linked to the species chosen. Enzyme assays (NRA) were shown to be a valuable tool for monitoring physiological status of planted trees subjected to environmental changes brought on by silvicultural practices. Lautenschlager, R. A. 1993. Effects of conifer release with herbicides on wildlife. (A review with an emphasis on Ontarios forests). Ontario Ministry of Natural Resources, Sault Ste. Marie, Ontario. VMAP Forest Research Information Paper No.111. 23 p. This paper reviews studies that have examined the effects of forest herbicide treatments on wildlife in Ontario and similar northern coniferous ecosystems. Since most of the research has been short-term, valid generalizations about the effects of conifer release with herbicides on specific wildlife species and/or groups (except moose) must be limited to 1 or 2 growing seasons after treatment. Moose browse and browse use reductions following conifer release with herbicides may last up to 4 years. However, early reductions in browse availability commonly found in treated areas may be offset by later additions. Such increases have been recorded 8 growing seasons after treatment. Although the data are limited, research indicates that if herbicide treatments reduce other (non-moose) local wildlife populations, those reductions likely last for only 1 to 2 years after treatment. However, more long-term research is needed. Conifer release is but one of several forest management tools designed to direct secondary succession and shape future forests. Managed forests, moreover, are only parts of larger ecosystems, which in turn are parts of the overall landscape. The effects of the composition of landscape mosaic on resident wildlife should be considered when conifer release is discussed. In Ontario, where harvesting has reduced the coniferous component in forests and landscapes, conifer release could help restore this component and increase both forest and wildlife diversity. . 1997. Effects of alternative conifer release treatments on small mammals in northwestern Ontario. Forestry Chronicle 73 (In press) . See Mammals Section.

38.

39.

21

Biodiversity, Conservation, and Habitat Restoration/Alteration 40. Lautenschlager, R. A., F. W. Bell, and R. G. Wagner. 1995. The Fallingsnow ecosystem project: comparing manual, mechanical, and aerial herbicide conifer release in northwestern Ontario. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R.E. Gaskin, and J.A. Zabkiewicz, (compilers). FRI Bulletin No. 192. pp. 262-264. This integrated study documents ecosystem responses and vegetation changes associated with alternative (aerial herbicides, manual release, mechanical release, control) conifer release treatments. Post-treatment control plots tended to have slightly larger, more diverse populations of the biotic components examined. Although data on reproductive fitness of censused animals have not been examined, few dramatic differences in species or densities (terrestrial gastropods, amphibians and reptiles, small mammals, or songbirds) were found among the alternative conifer release treatments studied. Lautenschlager, R. A., and M. L. McCormack, Jr. 1989. Herbicide release may increase species diversity. Forest and wildlife management in New England - What can we afford? R. D. Briggs, W. B. Krohn, J. G. Trial, W. D. Ostrofsky, and D. B. Field (editors). University of Maine, Orono, Maine: Cooperative Forestry Research Unit. The effect of the application of three herbicides commonly used for conifer release in Maine on the number of plant species occupying treated and control areas on a somewhat poorly drained and a well-drained site was examined. The total number of plant species and/or species groups growing on Triclopyr (applied at 1 and 2 qts/acre), Glyphosate (at 1 and 2 qts/acre) and Hexazinone (at 4 qts/acre) treated plots one and two years after their application was determined. Results indicated that these chemicals, at these rates, do not adversely affect, and may actually increase the total number of plant species occupying regenerating clearcut stands after their application. It was concluded that conifer release with herbicides does not reduce the number of plant species found in treated areas one and two years after their application. Lautenschlager, R. A., T. P. Sullivan, and R. A. Wagner. 1995. Using herbicides for wildlife management in northern ecosystems. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R. E. Gaskin, and J. A. Zabkiewicz (compilers). FRI Bulletin No. 192. pp. 152-54. The authors examine the effects of herbicides on wildlife habitat (based on the literature and their experience) in northern ecosystems and suggest specific vegetation management approaches using herbicides that have or could improve habitat for a variety of wildlife species and groups. By choosing appropriately (active ingredient, time, application technique), herbicides can: (1) reduce densities of invading non-native plants (restoring native populations and associated wildlife); (2) create snags, dead and down woody material, and drumming logs in early or later successional stands (providing old growth characteristics); (3) create small, intermediate, or large early successional openings within older vegetation types; (4) change shrub dominated areas to earlier successional grassy, or herb/grass dominated communities; (5) favor male aspen clones; (6) release patches or expanses of conifers; and (7) keep woody and herbaceous browse within reach of browsing animals. Linz, G. M., D. L. Bergman, and W. J. Bleier. 1992. Progress on managing cattail marshes with Rodeo herbicide to disperse roosting blackbirds. Proceedings of the 15th Vertebrate Pest Conference J.E. Borrecco & R.E. Marsh (editors), University of California, Davis. pp. 56-61. See Birds Section. Linz, G. M., D. L. Bergman, D. C. Blixt, and W. J. Bleier. 1994. Response of black terns (Chlidonias niger) to glyphosate-induced habitat alterations on wetlands. Colonial Waterbirds 17, no. 2 : 160-167 . See Birds Section. Linz, G. M., D. L. Bergman, H. J. Homan, and W. J. Bleier. 1995. Effects of herbicide-induced habitat alterations on blackbird damage to sunflower. Crop Protection 14, no. 8 : 625-29 . See Birds Section.

41.

42.

43.

44.

45.

22

Biodiversity, Conservation, and Habitat Restoration/Alteration 46. Linz, G. M., D. C. Blixt, D. L. Bergman, and W. J. Bleier. 1996. Response of ducks to glyphosateinduced habitat alterations in wetlands. Wetlands 16, no. 1 : 38-44 . See Birds Section. . 1996. Responses of red-winged blackbirds, yellow-headed blackbirds and marsh wrens to glyphosate-induced alterations in cattail density. Journal of Field Ornithology 67, no. 1 : 167-76 . See Birds Section. Lloyd, R. A. 1990. Impact on vegetation after operational Vision treatment at varying rates in the Skeena region. B.C. Ministry of Environment, Fish and Wildlife Branch, Smithers, B.C. See Mammals Section. Malecki, R. A., and T. J. Rawinski. 1985. New methods for controlling purple loosestrife (Lythrum alicaria). New York Fish and Game Journal 32: 9-19. New methods for controlling purple loosestrife were investigated ... in central New York during 1978-1980. Neither late-summer cutting nor partial flooding had more than a seasonal impact on the plant, but cutting followed by flooding in the same year may result in significant stress. ... The herbicide glyphosate was extremely effective (100 percent control) on test plots when applied at a rate of 1.7 kg/ha in mid-August. Viability of seeds from treated plants was sufficient for concern. Sowing Japanese millet on exposed moist-soil sites at a rate of 34 kg/ha following a drawdown resulted in successful control of loosestrife on those areas as well as production of a desirable waterfowl food. Marrs, R. H., A. J. Frost, R. A. Plant, and P. Lunnis. 1993. Determination of buffer zones to protect seedlings of non-target plants from the effects of glyphosate spray drift. Agriculture, Ecosystems, and Environment 45, no. 3-4 : 283-93 . There is an increasing need to protect semi-natural vegetation from the potential effects of herbicide drift. One way to protect sensitive sites is to surround them with a no-spray buffer zone. Earlier estimates of buffer zone size based on bioassay experiments with established perennials suggested zones needed to be 6-10 m wide. In this paper four bioassay experiments are reported, where seedlings grown in trays were exposed downwind of glyphosate applications and taken to a glasshouse for assessment. Three experiments were done with Lychnis flos-cuculi seedlings including one with different surrounding grass structures, and Experiment 4 tested the response of 15 species typical of semi-natural vegetation. The mortality of Lychnis flos-cuculi varied between experiments and appeared more or less unaffected by grassland structure except immediately downwind of the sprayer. The multi-species experiment indicated a wide sensitivity to spray drift, and one species was affected between 15 and 20 m downwind. Thus, seedlings of some species were affected at greater distances than established plants, indicating either greater capture of drift or a greater sensitivity. On sites where seedling establishment is an important mechanism for community regeneration, buffer zones may need to be 20 m wide. May, T. A., D. E. May, and M. L. McCormack, Jr. 1982. Plant response to herbicides in clearcut conifer strips. Proceedings of the 36th Annual Meeting of the Northeastern Weed Science Society, 36:205-8. Plant species composition and abundance were measured in regenerating (4-5 years) clearcut strips in Maine sprucefir forests prior to, one year after, and three years after application of varying rates of glyphosate, glyphosate in combination with 2,4-D, and triclopyr. Changes in plant species indices were used to categorize plant response to the herbicides used. The implications of these changes to wildlife use of herbicide-modified clearcut strips are discussed. McComb, W. C., L. R. Curtis, K. Bentson, M. Newton, and C. L. Chambers. 1990. Toxicity analyses of glyphosate herbicide on terrestrial vertebrates of the Oregon coast range., Final Report. U.S.D.A. Forest Service, Oregon State University. Final Report. 30 p. See Mammals Section.

47.

48.

49.

50.

51.

52.

23

Biodiversity, Conservation, and Habitat Restoration/Alteration 53. . 1997. Toxicity of glyphosate herbicide to terrestrial mammals and amphibians of the Oregon Coast Range. (Submitted). We compared intraperitoneal LD50s for glyphosate in 4 wild mammal and 5 wild amphibian species with white lab mice (Mus musculus). White mice seemed to be adequate models for 7 of the 9 species; data were insufficient to allow comparisons with the other 2 species. LD50s were >800 mg/kg, which was estimated to be 165 times higher than a dose likely to be received by these species after field application of glyphosate. Rough-skin newts (Taricha granulosa) and Townsend's chipmunks (Tamias townsendii) given sub-lethal doses of glyphosate were marked with radio transmitters and released into unsprayed habitat. These 2 species did not have detectably lower survival nor different movement patterns than individuals given a control substance. McMillan, L. M., J. F. Connor, H. R. Timmermann, J. G. McNicol, and C. S. Krishka. 1990. Small mammal and lesser vegetation response to glyphosate tending in North Central Ontario. Ontario Ministry of Natural Resources, Northwestern Ontario Forest Technology Development Unit, Ontario. Technical Report No. 57. 28 p. See Mammals Section. Messersmith, C. G., Z. J. Woznica, R. G. Lym, and K. B. Thorsness. 1994. Evaluation of aquatic herbicides for managing blackbird roosting habitat. USDA Animal and Plant Health Inspection Service , North Dakota State University, Fargo, ND. Final Report. Blackbirds prefer cattail marshes as roosting sites and preferentially feed on nearby sunflower fields thereby inflicting great losses to farmers. Blackbirds can be dispersed by controlling cattails so feeding damage to individual farms is minimized. We evaluated cattail control by glyphosate at four rates and four cattail growth stages, with four adjuvants, and with or without diammonium sulfate. Glyphosate ground-applied at 1.7 kg/ha (1.5 qt RODEO/A) or greater generally controlled cattail similarly regardless of growth stage, but late July (flowering stage) tended to be the optimal time. Glyphosate helicopter-applied at 2.2 kg/ha (2 qt RODEO/A) tended to be more effective than at 1.7 kg/ha, although streaking during application may have accounted for reduced control when the lower rates were applied. Cattail control in the field with glyphosate at 1.7 kg/ha or greater requires a surfactant but generally was not affected by the specific adjuvant or diammonium sulfate added in the spray solution. The effects of plant characteristics and adjuvants on glyphosate retention, absorption and translocation were evaluated using glyphosate plus dye or 14C-glyphosate applied to greenhousegrown cattail. 14C-glyphosate absorption reached near maximum level within 24 h after treatment and was highest from the lower leaf surface, near the leaf tip, and with the youngest leaves. Not all surfactants enhanced glyphosate absorption by cattail compared to not using a surfactant, but the most effective surfactants enhanced absorption two- to three-fold, and cationic surfactants generally were more effective than nonionic surfactants. Similarly, glyphosate retention with several surfactants varied by magnitudes as great as absorption. Among commercially available adjuvants, Kinetic (TM) at 0.5% (v/v) applied alone and R-11 (TM) adjuvant at 0.5% (v/v) plus Chem-Trol (TM) polyvinyl polymer at 1% (v/v) should be the most effective adjuvants for cattail control. Miller, K. V., and J. S. Witt. 1990. Herbicide-Wildlife interactions. Forest Vegetation Management Workshop Proceedings C.W. Murdoch (editor). College of Forest Resources, University of Maine. pp. 20-24. In summary, the use of herbicides for vegetation management is not as detrimental to wildlife habitat as once assumed. In fact, the use of herbicides often may be superior to other vegetation management methods in terms of their effects on wildlife. Herbicides should be regarded as a tool for use by forest and wildlife managers. It is important to remember, however, that herbicides are not the only tools available. All man-caused activities in forested systems can impact wildlife habitat, and it is important to understand the results of these activities. The selectivity of the newer generations of herbicides suggest exciting possibilities regarding their potential uses. The vastly differing plant communities that develop following the use of different chemicals necessitates additional study of their impacts on wildlife species. However, provided suitable data is obtained, it may one day be possible to tailor both types and rates of herbicide application to not only achieve vegetation management goals, but also selectively enhance or reduce populations of local wildlife species. 24

54.

55.

56.

Biodiversity, Conservation, and Habitat Restoration/Alteration 57. . 1991. Impacts of forestry herbicides on wildlife. Proceedings of the Sixth Biennial Southern Silvicultural Research Conference, S.S. Coleman, and D.G. Neary (editors). Asheville, North Carolina : Southeastern Forest Experiment Station. Vol. 2: 795-800. Concern over the widespread use of herbicides arises from two factors: potential direct toxicity to wildlife and indirect effects on wildlife habitat. From the results of our studies and other published literature, it is apparent that wildlife habitat is not adversely affected by herbicidal site preparation, except perhaps during the first growing season. Herbicide use at times may be superior to other vegetation management methods in terms of their effects on wildlife. However, the vastly differing plant communities that develop following the use of different chemicals necessitates additional study of their impacts on wild species. Milton, G. R., and J. Towers. 1990. Relationships of songbirds and small mammals to habitat features on plantation and natural regeneration sites. Canadian Institute of Forestry, St. Mary's River Forestry-Wildlife Project. Report No. 7. 57 p. See Mammals Section. Moola, F. M., A. U. Mallik, and R. A. Lautenschlager. 1997. Effects of conifer release treatments on blueberry production in northwestern Ontario. Canadian Journal of Forest Research (In press). Berry production and vegetative recovery of lowbush blueberry (Vaccinium angustifolium) and velvet leaf blueberry ( Vaccinium myrtilloides) were documented for three growing seasons (1994, 1995, 1996) after manual brushsaw cutting, single operational and multiple non-operational Vision herbicide (a.i., glyphosate), and control treatments in a young jack pine (Pinus banksiana) plantation near Atikokan, northwestern Ontario. Total berry number per hectare (both species combined) was significantly reduced by 34% and 58% following operational single Vision treatment in 1995 and 1996, respectively. However, V. angustifolium was more sensitive to the herbicide treatments than V. myrtilloides. Its berry yield was consistently lower in both the single and multiple Vision treatments in all three years of the study. Berry yield of V. myrtilloides in 1994, 1995, and 1996 was significantly reduced by 87, 100 and 100%, respectively, only by the multiple Vision treatment. The pubescent foliage of V. myrtilloides might have hindered absorption of glyphosate and thus decreased its sensitivity to single Vision treatment. Discriminant analysis of 12 vegetative parameters (height, percent cover, number of unaffected, partially defoliated, severely defoliated and dead stems, dry weight of live and dead stems, dry weight of rhizomes and leaves, leaf area and leaf weight) of the two Vaccinium spp. suggests that the recovery of both species of blueberry was significantly different in Vision treated plots compared to the brushsaw and control plots. Both species exhibited greater stem defoliation and mortality, lower percent cover and lower leaf area in the single and multiple Vision treatments. Reduced total blueberry yield associated with the single Vision treatment could have shortterm effects on resident wildlife, including black bear (Ursus americanus) in treated areas. Conversely, since manual brushsaw cutting of competing vegetation stimulates both vegetative regeneration and fruit yield of Vaccinium spp., it provides a better vegetation management alternative that combines conifer release and enhanced berry production. Morrison, M. L., and E. C. Meslow. 1983. Impacts of forest herbicides on wildlife: Toxicity and habitat alteration. Transactions of the 48th North American Wildlife Conference. pp. 175-185. See Mammals Section. Mullahey, J. J., J. A. Cornell, and D. L. Colvin. 1993. Tropical soda apple (Solanum viarum) control. Weed Technology 7 : 723-27. Hexazinone (1.12 kg ai/ha), triclopyr (1.12 kg ai/ha), metsulfuron (0.008 kg a.i./ha), dichlorprop + 2,4-D, glyphosate (2.8%), and triclopyr (2%) + diesel oil (98%), applied as a broadcast or spot (individual plant) treatment, were evaluated over two years in south Florida for tropical soda apple (TSA) control and their effects on grass ground cover. For broadcast treatments, triclopyr (98%) and hexazinone (93%), had significantly (P < 0.05) higher percent control of marked TSA plants 90 d after herbicide application. However, triclopyr (99%) had significantly higher grass ground cover than 25

58.

59.

60.

61.

Biodiversity, Conservation, and Habitat Restoration/Alteration hexazinone (78%). Hexazinone severely damaged Pangola digitgrass, but had no effect on bihiagrass. For spot treatments, dichlorprop + 2,4-D (100%) had the highest percent total control of TSA and least effect on grass ground cover (96%) 90 d after herbicide application, followed by glyphosate (96% control) and triclopyr + diesel oil (95% control). Based on acceptable (>90%) TSA control and grass ground cover, triclopyr broadcast or dichlorprop + 2,4-D spot provided the greatest control. With either application method, repeated herbicide applications will be necessary to eliminate TSA because of rapid seedling emergence following control of existing plants. 62. Neary, D. G., and J. L. Michael. 1995. The role of herbicides in protecting long-term sustainability and water quality in forest ecosystems. Second International Conference of Forest Vegetation Management, Rotorua, New Zealand. R.E. Gaskin and J.A. Zabkiewicz (compilers). FRI Bulletin No. 192, pp. 161-63. See Water Quality Section. Neary, D. G., J. E. Smith, B. F. Swindel, and K. V. Miller. 1990. Effects of forestry herbicides on plant species diversity. Proceedings Southern Weed Science Society Risk/Benefit: A way of life 43rd Annual Meeting , pp. 266-72. Two sites in the north Florida flatwoods were treated with sulfometuron methyl, glyphosate, and triclopyr for site preparation. Plant diversity changes were assessed by the line transect method on both an intensive (annual applications) and less intensive (two applications) site. On the intensivelytreated site, plant species were reduced from a range of 4-16 to 0-7 per transect. Also, 18% of the transects were bare or contained only pine trees. On the less-intensively treated site, the plant species distribution shifted only slightly, but total numbers were reduced. Further research is needed to quantify the effect of operational conditions and other herbicides on plant species diversity. Newton, M., E. C. Cole, M. L. McCormack, Jr. and D. E. White. 1992. Young spruce-fir forests released by herbicides II. Conifer response to residual hardwoods and overstocking. Northern Journal of Applied Forestry 9, no. 4 : 130-135 . Responses of conifers and other vegetation to 9 aerial herbicide treatments were evaluated in a replicated conifer-release experiment in a 7-year-old spruce-fir clearcut in central Maine. Development of the naturally regenerated conifers was inversely related to residual hardwood cover and conifer stocking during the 9 years after treatment. All herbicides and rates of application initially reduced hardwood cover by 50% or more. Cover more than 1.5 m tall was nearly eliminated by several treatments, and release from overtopping provided long-term increase in conifer growth. Untreated hardwoods severely reduced dominance and stocking of conifers by age 16. Spruce heights and diameters were less affected by hardwood competition than were those of fir; current growth in released fir was greater than that in released spruce; but unreleased spruce, where present, grew faster than unreleased fir. Conifer growth was affected both by residual hardwood overtopping and by number of conifers within 0.91 m. Newton, M., E. C. Cole, D. E. White, and M. L. McCormack, Jr. 1992. Young spruce-fir forests released by herbicides I. Response of hardwoods and shrubs. Northern Journal of Applied Forestry 9, no. 4 : 126-30 . Responses of shrubs and hardwoods to 9 aerial herbicide treatments were evaluated in a replicated conifer-release experiment in a 7-year-old spruce-fir clearcut in west-central Maine. All herbicides and rates of application reduced hardwood and shrub cover by 50% or more in year 9, 2 years after treatment. Cover more than 1.5 m tall was nearly eliminated by treatments with triclopyr amine (Garlon 3A, glyphosate (Roundup), or a high rate of 2,4,5-T. Phenoxy herbicides (2,4-D and 2,4,5-T) led to short-term reductions in birches, maples, aspen, and raspberry, and little change in willows. Pin cherry was also controlled by these treatments but died out before year 16, regardless of whether it was sprayed. Untreated controls increased in total cover by about 50% between years 7 and 9 and also in abundance of cover >1.5 m tall. Much cover less than 1.5 m tall was left by all treatments other than triclopyr at a high rate. By the 16th year, major differences in height and cover still existed between all treatments and the controls.

63.

64.

65.

26

Biodiversity, Conservation, and Habitat Restoration/Alteration 66. Newton, M., R. Willis, J. Walsh, E. Cole, and S. Chan. 1996. Enhancing riparian habitat for fish, wildlife, and timber in managed forests. Weed Technology 10 : 429-38 . The productivity of riparian sites in managed forests can be focused to provide productive fish and wildlife habitat while yielding most of its productive capacity for other than amenity values. Establishment of habitat protection goals and measures of achievement permit flexible approaches for meeting them. Once the protection standards are set, intensive management of the woody cover is logically dependent on minimum disturbance methods, in general, for both vegetation management and harvest. Several currently registered chemical products and non-chemical methods are helpful and safe in achieving both yield and protection goals. Nickerson, N. H. 1992. Impacts of vegetation management techniques on wetlands in utility rights-of-way in Massachusetts. Journal of Arboriculture 18: 102-6. This project compared five rights-of-way (R/W) treatments (hand cutting, mowing, cut stump treatment w/herbicides, basal spraying w/herbicide, and foliar application of herbicides) to determine their impacts on wetlands on utility R/W. The conclusion reached was that there was no significant impact to wetlands from any of the vegetation management techniques used on utility R/Ws in Massachusetts. Mechanical treatments resulted in higher impacts on the cover value for wildlife than those involving herbicides. Residue from petroleum products (baroil and hydraulic fluid) were recovered on the leaf litter from mechanically treated sites. No herbicide residues were recovered from herbicide-treated sites. Nuzzo, V. A. 1991. Experimental control of garlic mustard [Alliaria petiolata (Bieb.) Cavara & Grande] in northern Illinois using fire, herbicide, and cutting. Natural Areas Journal 11, no. 3:158-67. Garlic mustard [Alliaria petiolata (Bieb.) Cavara & Grande] is a naturalized European obligate biennial herb that invades forest communities in the midwestern and northeastern United States and southeastern Canada. Three potential control methods (prescribed fire, 3% v:v glyphosate, cutting) were tested in a densely infested oak forest in northern Illinois. Spring treatment with either glyphosate or mid-intensity fire reduced garlic mustard adult (rosette) density and seedling frequency, while fall treatment decreased rosette density only. Low-intensity fire did not affect garlic mustard presence. All successful treatments produced cascading effects observable one to two years after treatment. Treating seedlings in spring reduced adult density the following year, and treating adults in either spring or fall resulted in lower seedling frequency two years later. Cutting flowering plants at ground level resulted in 99% mortality and reduced seed production to virtually zero; cutting at 10 cm above ground level produced 71% mortality and reduced total seed production by 98%. Recommended management is to prevent seed production until the seed bank is exhausted by repeated applications of fire, herbicide, or cutting on an annual basis until garlic mustard is absent from the site for a minimum of three years. Olaleye, V. F., and O. A. Akinyemiju. 1996. Effects of glyphosate (N-(phosphonomethyl) glycine) application to control Eichhornia crassipes Mart. on fish composition and abundance in Abiala Creek, Niger Delta, Nigeria. Journal of Environmental Management 47, no. 2 : 115-22 . See Fish Section . Petrie, G. A. 1995. Effects of chemicals on ascospore production by Leptosphaeria maculans on blackleg-infected canola stubble in Saskatchewan. Canadian Plant Disease Survey 75, no. 1:45-50. Chemical treatment of blackleg-infected stubble was often very effective in inhibiting or eliminating the production of ascospores by Leptosphaeria maculans. Among the most effective fungicides were the ergosterol biosynthesis inhibitors, nuarimol, propiconazole, and triadimenol, which at 0.1 g a.i./L reduced spore numbers by 99-100%. Prochloraz, imazalil, chlorothalonil, and benomyl also significantly reduced ascospore numbers at 0.1 g a.i./L. At 10 g a.i./L, the herbicide glyphosate completely prevented ascospore formation, whereas trifluralin increased sporulation. A role for chemical treatment of stubble for control of blackleg on canola is indicated in light of the trend toward minimum tillage practices by western Canadian producers.

67.

68.

69.

70.

27

Biodiversity, Conservation, and Habitat Restoration/Alteration 71. Pojar, R. 1990. The effects of operational application of the herbicide glyphosate on target brush species, selected browse species and conifer crop trees. Two years after application. Ministry of Forests Progress Report, B.C. Vol. 3. p. 49. See Mammals Section. Pywell, R. F., N. R. Webb, and P. D. Putwain. 1995. A comparison of techniques for restoring heathland on abandoned farmland. Journal of Applied Ecology 32 : 400-411 . 1. Recent changes in agricultural policies have reduced the extent of cultivated farmland. This has provided opportunities to restore heathland vegetation on lowland sites where it once occurred. 2. Between December 1988 and April 1990 large-scale replicated experiments were established on abandoned farmland in southern Britain to compare the effectiveness of four treatments for heathland restoration: (i) the application of herbicide; (ii) the addition of harvested heather shoots; (iii) the addition of heathland topsoil; and (iv) the translocation of heathland turves. 3. The number of seedlings of heathland plant species on each treatment was counted in December 1990 and 1991, and the shoot frequency of these species was recorded in January 1993. 4. The grassland soil had a significantly higher pH and contained greater concentrations of extractable phosphorus and exchangeable calcium than that of the adjacent heathland. Despite this, the controls showed that there was some natural regeneration of heathers within the grassland. 5. Herbicide treatment inhibited the regeneration of heathland plants. Cultivation followed by the application of harvested heather shoots increased the number of seedlings of heathland plant species, but some key species were missing. All the components of the heathland plant community occurred in greater numbers on the plots where heathland topsoil had been applied, and on the parts of transferred heathland turves which had died from drought. 6. The large-scale translocation of heathland turf appeared to be feasible and instantly recreated the mature heathland plant community. However, some changes in the plant community occurred which probably resulted from differences in soil drainage characteristics between the donor and recipient sites. Of the different sources of heathland plant propagules, harvested heather shoots were a renewable resource, whereas the collection of heathland topsoil and turves involved the destruction of existing heathland. Randall, J. M. 1996. Weed control for the preservation of biological diversity. Weed Technology 10: 370-383. Invasions by non-native plants threaten the preservation of many plant and animal species and communities throughout North America. These pest species compete with and displace native plants and animals and may substantially alter ecosystem functions (e.g., fire occurrence and frequency, nutrient cycling). Awareness of these threats among wildland managers has greatly increased in the last decade. In a recent poll of National Park superintendents, 61% of 246 respondents indicated nonnative plant invasions were moderate or major problems at their parks. Likewise, over 60% of Nature Conservancy stewards nationwide polled in 1992 indicated weeds were among their top 10 management problems, listing nearly 200 problem species. Over 12% indicated weeds were their worst problem. Weed control programs are now in place in wildlands across the continent, employing techniques ranging from manual removal, mechanical methods, prescribed fire, judicious use of herbicides, the release of biological control agents, and encouragement of native competitors. The most successful endeavors follow an adaptive management strategy in which plans based on the goals of the preserve are deveolped, weeds that interfere with those goals are identified and prioritized, and control measures are selected and implemented where appropriate. Emphasis is placed on preventing new weeds from becoming established and on early detection and elimination of incipient infestations. Managers must focus on the vegetation or community desired in place of the weeds and periodically reevaluate whether their programs are moving them toward this objective. Control of weeds in wildlands poses unusual problems not ordinarily met in other systems which offer challenging research opportunities for weed scientists and ecologists. Raymond, K. S., F. A. Servello, B. Griffith, and W. E. Eschholz. 1996. Winter foraging ecology of moose on glyphosate-treated clearcuts in Maine. Journal of Wildlife Management 60, no. 4 : 753-63 . See Mammals Section.

72.

73.

74.

28

Biodiversity, Conservation, and Habitat Restoration/Alteration 75. Riemer, D. N. 1976. Long-term effects of glyphosate applications to phragmites. Journal of Aquatic Plant Management 11: 39-43. N-(phosphonomethyl) glycine (glyphosate) was applied to phragmites (Phragmites communis Trin.) and the plots were observed for four growing seasons. The optimum rate of application appears to be in the range of 4 to 6 lb ae/A. Additional surfactant is effective only when the herbicide is applied at low rates. Within the range of 20 gpa to 80 gpa the effects of spray volume are minimal and can be ignored except with very low rates of herbicide. The application of glyphosate to phragmites for two successive years is a very effective means of control, even at rates as low as 2 lb ae/A each year. Rodgers, R. D. 1983. Reducing wildlife losses to tillage in fallow wheat fields. Wildlife Society Bulletin 11: 31-38. See Birds Section. Roundup. 1991. From Prairie to Pompeii., Monsanto Company. Glyphosate-based herbicides, like Roundup, Rodeo, Azural and Accord, are ideally suited to land management and wildlife restoration projects worldwide. The effectiveness and desirable characteristics of glyphosate are important reasons why the organizations mentioned in this booklet chose one of these herbicides for their delicate restoration projects. Used according to label directions, Roundup and other herbicides can eradicate weeds and create a beneficial environment for the regrowth of desirable vegetation without fear of harm to wildlife or the environment. Glyphosate-based herbicides by Monsanto have been thoroughly reviewed and registered by the U.S. Environmental Protection Agency and by regulatory agencies in more than 100 countries around the world. Roy, D. N., S. K. Konar, S. Banerjee, D. A. Charles, D. G. Thompson, and R. Prasad. 1989. Uptake and persistence of the herbicide glyphosate (VisionR) in fruit of wild blueberry and red raspberry. Canadian Journal of Forest Research 19: 842-47. See Plant and Soil Residues Section. Runciman, J. B., and T. P. Sullivan. 1996. Influence of alternative conifer release treatments on habitat structure and small mammal populations in south central British Columbia. Canadian Journal of Forest Research 26, no. 11: 2023-34. See Mammals Section. Sale, J. S. P., and P. M. Tabbush, and others. 1986. The use of herbicides in the forest - 1986. Edinburgh. Forestry Commission booklet 51. 122 pp. (United Kingdom). Santillo, D. J. 1994. Observations on moose, Alces alces, habitat and use on herbicide-treated clearcuts in Maine. Canadian Field-Naturalist 108, no. 1 : 22-25 . See Mammals Section. Santillo, D. J., D. M. Leslie, Jr. , and P. W. Brown. 1989. Responses of small mammals and habitat to glyphosate application on clearcuts. Journal of Wildlife Management 53: 164-72. See Mammals Section. Servello, F. A., B. Griffith, K. S. Raymond, and W. E. Eschholz. 1995. Effects of glyphosate on winter habitat of moose in Maine. College of Natural Resources, Forestry and Agriculture, Maine Agricultural and Forest Experiment Station, University of Maine, Orono, Maine, Cooperative Forestry Research Unit . CFRU Research Bulletin 10. Miscellaneous Report 395. 19 p. See Mammals Section.

76.

77.

78.

79.

80.

81.

82.

83.

29

Biodiversity, Conservation, and Habitat Restoration/Alteration 84. Simenstad, C. A., J. R. Cordell, L. Tear, L. A. Weitkamp, F. L. Paveglio, K. M. Kilbridge, K. L. Fresh, and C. E. Grue. 1996. Use of Rodeo and X-77 spreader to control smooth cordgrass (Spartina alterniflora) in a southwestern Washington estuary: 2. Effects on benthic microflora and invertebrates. Environmental Toxicology and Chemistry 15, no. 6 : 969-78 . See Aquatic Invertebrates and Algae Section. Solberg, K. L., and K. F. Higgins. 1993. Effects of glyphosate herbicide on cattails, invertebrates, and waterfowl in South Dakota wetlands. Wildlife Society Bulletin 21 : 299-307 . See Birds Section. Stockard, J. D. 1996. Restoration of Wingham brush 1980-1996. Eleventh Australian Weeds Conference Proceedings. pp. 432-36. The remnant rainforest at Wingham Brush was smothered in weeds and dissected with roads and tracks. Australia's first attempt to restore a rainforest started at Wingham in 1980. The impacts of flying-foxes, climate and humans are discussed in relation to retention of the forest. The development of glyphosate application techniques for aggressive exotic vegetation formed the basis of the programs success. The major goal in rainforest restoration is gaining a weed-free canopy and closing it. Continual maintenance weeding is required to ensure native succession and the forests survival. (Australia) . Sturko, A. 1996. Management of weeds on forest rangelands. Integrated Forest Vegetation Management: Options and Applications. Proceedings of the fifth B.C. Forest Vegetation Management Workshop. P. G. Comeau, G. J. Harper, M. E. Blache, J. O. Boateng, and L. A. Gilkeson (compilers). pp. 75-81. Many non-native plant species have been intentionally or accidentally introduced into British Columbia. Several of these weed species are prolific reproducers and aggressive competitors and have expanded their distribution dramatically. These weeds now pose a serious threat to significant areas of rangelands throughout British Columbia. Over 90 000 hectares are infested with noxious weeds, which have reduced levels of forage for wildlife and livestock and have negatively impacted native plant communities. While the primary threat is from diffuse and spotted knapweed (Centaurea diffusa Lam. and C. maculosa Lam.), other weed species such as Canada thistle (Cirsium arvense L. Scop.), hounds-tongue (Cynoglossum officinale L.), dalmatian toadflax (Linaria dalmatica (L.) Miller), leafy spurge (Euphorbia esula L.) and sulphur cinquefoil (Potentilla recta L.) pose significant problems. The Ministry of Forests range program integrates the use of chemical, cultural, and biological control treatments to meet their weed management objectives. These objectives include: protecting presently non-infested areas; preventing the establishment of new weed species presently not found on Crown range in British Columbia; and introducing biological control agents to reduce target weed species density, to a level where they no longer pose an ecological threat to resources or a socio-economic threat to the public using those resources. Disturbed areas are revegetated to minimize the opportunity for new weed infestations. Small weed infestations and transportation corridors are targeted for chemical and cultural treatments and large infestations of weeds are targeted for biological control as biocontrol agents become available. Sullivan, T. P. 1994. Influence of herbicide-induced habitat alteration on vegetation and snowshoe hare populations in sub-boreal spruce forest. Journal of Applied Ecology 31 : 71730 . See Mammals Section. Sullivan, T. P., and J. O. Boateng. 1996. Comparison of small-mammal community responses to broadcast burning and herbicide application in cutover forest habitats. Canadian Journal of Forest Research 26 : 462-73 . See Mammals Section. Sullivan, T. P., R. A. Lautenschlager, and R. G. Wagner. 1995. Changes in diversity of plant and small mammal communities after herbicide application in sub-boreal spruce forest. Second

85.

86.

87.

88.

89.

90.

30

Biodiversity, Conservation, and Habitat Restoration/Alteration International Conference on Forest Vegetation Management. Rotorua, New Zealand. R.E. Gaskin and J.A. Zabkiewicz (compilers). FRI Bulletin No. 192. pp. 143-45. See Mammals Section. 91. . 1996. Influence of glyphosate on vegetation dynamics in different successional stages of sub-boreal spruce forest. Weed Technology 10 : 439-46 . This study was designed to evaluate changes in plant biomass, species richness, and diversity after application of glyphosate herbicide in several successional stages of sub-boreal spruce forest near Prince George, British Columbia, Canada. Vegetation was sampled in replicate control (reference) and treatment blocks of herb, shrub, and shrub-tree stages of cutover forest habitats. Volume of herb layers declined temporarily in the first post-treatment year. Shrub layers were reduced in herb and shrub stages, and shrubs and trees were reduced temporarily in the shrub-tree stage. Species richness of herbs and shrubs was similar in control and treatment blocks in the herb successional stage, but shrub richness declined on the treatment in the shrub stage. There were no consistent differences in numbers of herb, shrub, or tree species between control and treatment blocks in the shrub-tree stage. Species diversity (Simpsons index and Shannon-Wiener function) of herbaceous plants was not affected by herbicide application in any of the successional stages, but diversity of shrubs was lower in treatment than control blocks in the herb and shrub stages. Diversity of trees was reduced on treatment blocks in the shrub-tree stage. Species abundance curves of overall plant communities showed little change in the herb stage, a decline in the first post-treatment year in the shrub and shrubtree stages, with similar patterns between control and treatment blocks in subsequent post-treatment years. The general lack of community-wide reductions in plant biomass and diversity, and the shortterm duration of specific changes, suggest that conifer release treatments of plantations have no substantial, incremental effects on wildlife habitat. Sullivan, T. P., C. Nowotny, R. A. Lautenschlager, and R. G. Wagner. 1997. Silvicultural use of herbicide in sub-boreal spruce forest: implications for small mammal population dynamics. (Submitted). See Mammals Section. Sullivan, T. P., and D. S. Sullivan. 1982. Responses of small-mammal populations to a forest herbicide application in a 20-year-old conifer plantation. Journal of Applied Ecology 19: 95106. See Mammals Section. Sullivan, T. P., D. S. Sullivan, E. J. Hogue, R. A. Lautenschlager, and R. G. Wagner. 1997. Population dynamics of small mammals in relation to habitat alteration in orchard agroecosystems: Compensatory responses in abundance and biomass. (Submitted). This study was designed to test the hypothesis that habitat alteration with intensive herbicide (glyphosate) treatment would adversely affect the small mammal community in apple orchards. Multiple applications of herbicide were applied to the total orchard floor in two experimental orchards and this same herbicide regime was applied in a 2-m-wide strip within tree rows only in a third orchard. Intensive sampling of montane vole (Microtus montanus), deer mouse (Peromyscus maniculatus), and northwestern chipmunk (Eutamias amoenus) populations was conducted in replicate control and treatment blocks of apple orchards at Summerland, British Columbia, Canada from 1983 to 1986. Average abundance of voles declined by 53% to 73% on treatment compared to control blocks after the first herbicide application. Vole populations were consistently reduced in response to the herbicide treatment, with average abundance ranging from 2.8 to 28.0 times higher on control than treatment blocks. Voles declined to, or near, extirpation in all orchards during the winter of 1985-86. Deer mouse and northwestern chipmunk populations were either significantly higher on treatment than control blocks or there was no difference in abundance after the start of the herbicide applications. Average abundance of deer mice ranged from 1.3 to 11.1 times higher, and that of chipmunks ranged from 1.8 to 13.3 times higher, on treatment than control blocks. The high numbers of deer mice and chipmunks on treatment blocks were composed mainly of resident animals. There were no significant differences in biomass of small mammals between control and treatment populations in summer and winter periods in 2 of 3 orchards. There appears to be a compensatory response in this small mammal community 31

92.

93.

94.

Biodiversity, Conservation, and Habitat Restoration/Alteration whereby deer mice and northwestern chipmunks have essentially replaced the montane vole on treatment blocks in an orchard agroecosystem. Quality of the altered habitats seemed sufficiently high to support these populations at comparable or higher levels than on controls. Species diversity of plants and animals in orchard agroecosystems should be investigated to determine their role as part of a mosaic of natural and agrarian habitats. 95. Sullivan, T. P., D. S. Sullivan, R. A. Lautenschlager, and R. G. Wagner. 1997. Long-term influence of glyphosate herbicide on demography and diversity of small mammal communities in coastal coniferous forest. Northwest Science 71, no. 1 : 6-17 . See Mammals Section. Sullivan, T. P., and L. W. Taylor. 1995. Wildlife habitat enhancement with Ezject in stand thinning of lodgepole pine. Second International Conference on Forest Vegetation Management, Rotorua, New Zealand. R.E. Gaskin and J.A. Zabkiewicz (compilers). FRI Bulletin No. 192. pp. 310-312 . This study is determining the responses of understorey vegetation and small mammal communities to stand thinning of lodgepole pine (Pinus contorta) by conventional cutting and by Ezject (injection herbicide capsules). These two techniques may have substantial effects on understorey vegetation, in terms of available light and debris on the forest floor. This will likely affect wildlife and also quality of forage for livestock using forests as summer range. Sullivan, T. P., R. G. Wagner, D. G. Pitt, R. A. Lautenschlager, and D. G. Chen. 1997. Changes in diversity of plant and small mammal communities after herbicide application in sub-boreal spruce forest. (Submitted) . This study tested the hypothesis that forest herbicide use would reduce species richness and diversity of plant and small mammal communities over a 5-year period in sub-boreal spruce forest near Prince George in central British Columbia, Canada. Four control (untreated) and four treatment sites of similar physiographic and vegetation characteristics were selected to quantify the response of shrub and herbaceous vegetation, and associated small mammal communities, that were treated with glyphosate herbicide for conifer release. Plant and small mammal communities were sampled during 1987 (pretreatment year), 1988-1989 (first and second posttreatment years), and in 1991-1992 (fourth and fifth posttreatment years). Herb crown volume recovered to pretreatment or control levels in the second posttreatment year. Shrub crown volume was reduced by 93% (P<0.0001) on all treatment sites after herbicide application. There was no difference (P=0.140) in number of herbaceous plant species between control and treatment sites. Plant species richness of shrubs, however, declined (P=0.001) on treatment compared with control sites. Total number of plant species generally increased with successional development through time on all sites. Species diversity of shrubs as measured by the Shannon-Wiener (S-W) index was lower (P=0.055) on treatment than control sites; but not when Simpsons index (P=0.120) was used. There was no difference (P>0.170) in herb diversity throughout the study. There were no differences in species richness (P=0.420) nor diversity (P>0.120) of small mammal communities inhabiting control and treatment sites. Herbicide application reduces vegetative biomass and shrub species diversity on a short-term basis, with no impact on herbaceous species richness or diversity in sub-boreal spruce forest. In general, diversity of plant and small mammal communities seemed to be maintained, and hence these treatment sites should not lower overall diversity of a forested landscape. Symondson, W. O. C., D. M. Glen, C. W. Wiltshire, C. J. Langdon, and J. E. Liddell. 1996. Effects of cultivation techniques and methods of straw disposal on predation by Pterostichus melanarius (Coleoptera: Carabidae) upon slugs (Gastropoda: Pulmonata) in an arable field. Journal of Applied Ecology 33: 741-53. 1. Interactions between the polyphagous carabid predator Pterostichus melanarius (Illiger) and slugs were investigated from July to September 1992, before and after harvesting a rape crop. The experimental site comprised a long-term field study of the effects of different forms of cultivation (ploughing vs. non-inversion tillage), and methods of straw disposal (baling vs. incorporation of chopped straw) upon invertebrate populations and crop yields. Direct-drilling was also included as a no-tillage base-line. 2. Beetles (total 2078) were collected by pitfall trapping twice weekly. Each beetle 32

96.

97.

98.

Biodiversity, Conservation, and Habitat Restoration/Alteration was dissected, and its crop contents weighed and tested by enzyme-linked immunosorbent assay (ELISA) to determine the concentration and quantity of slug haemolymph it contained. Slugs [Deroceras reticulatum (Muller) and Arion intermedius Normand] were extracted from soil samples by gradual flooding, to estimate both numbers and biomass. 3. Significantly more P. melanarius were trapped in direct-drilled plots than in the tilled treatments. Within the tilled treatments, greater numbers of beetles were trapped where straw was incorporated by non-inversion tillage. 4. Crop weights were significantly greater in beetles from direct-drilled plots than in those from tilled treatments, as were both the concentrations and quantities of slug haemolymph they contained. Overall, 84% of beetles contained slug remains. 5. Greatest concentrations and quantities of slug remains were detected prior to the disposal of rape residues at the end of July, by baling or shallow incorporation in the soil. Cultivation had both short- and long-term effects upon the proportion of the diet of the beetles that was slugs. 6. Slug biomass declined following disposal of rape residues and it was only after this time that significant treatment differences emerged. 7. Positive relationships were found between the biomass of slugs in the soil and numbers of beetles trapped, the proportion of the beetles diet that was slugs and the quantities of slug haemolymph in beetle crops. 8. Our results strongly suggest aggregation of P. melanarius to areas of high slug biomass in the soil and preferential feeding in such areas upon slugs. As this carabid is probably the commonest large predatory beetle in arable crops in Britain, these results clearly identify P. melanarius as a potentially important slug control agent. 99. Tanner, G. W., J. M. Wood, and S. A. Jones. 1992. Cogongrass (Imperata cylindrica) control with glyphosate. Florida Scientist 55, no. 2 : 112-15 . A stand of cogongrass was burned in February 1983, and the regrowth sprayed with 0.5, 1.0, and 2.0% solutions (c/c) of glyphosate in either April, August, or November of the same year. Plant density was estimated immediately before each treatment and again in July 1984 and 1985. Plant control with the 2% glyphosate treatment differed significantly (p lt 0.05) from the 0.5 and 1.0% solution treatments only after the November application date. Glyphosate as a 2% solution appeared to control cogongrass satisfactorily for at least 2 yr (91% mortality), while control after the August application never exceeding 40%. Controlling cogongrass in forests will be required to effectuate adequate natural regeneration of pines or survival of planted seedlings and to maintain natural diversity of plant species. Tenuta, M., and R. G. Beauchamp. 1996. Denitrification following herbicide application to a grass sward. Canadian Journal of Soil Science 76, no. 1 : 15-22 . The application of the herbicide Roundup (glyphosate), and subsequent death of a predominantly bromegrass (Bromus inermis Leyss.) and blue grass (Poa pratensis L.) sward, resulted in a 20- to 30-fold increase in denitrification rate 14 and 49 d after application compared to herbicideuntreated and fallowed soil treatments. The regulation of denitrification by O2, carbon and NO-3availability was assessed by measurement of various soil variables. The regulation of denitrification by C and NO-3- availabilities was further studied in a laboratory experiment in which denitrification was measured following NO-3- and glucose-C addition to soil from the field treatments. Elevated denitrification in the herbicide-treated soil was attributed to increased soil moisture and NO-3- contents resulting from the death of vegetation. The death of the grass sward did not increase available C to denitrifiers, whereas the absence of vegetation in the fallowed soil 1 yr following herbicide application reduced available C. This study indicates that herbicide application to a grass sward increases denitrification and hence may contribute to greater nitrous oxide emission and N loss from soil. Timmerman, H. R., J. G. McNicol, and C. S. Krishka. 1986. Impact of glyphosate on wildlife habitat: A three year study. Ministry of Natural Resources, Wildlife and Forest Resources Branch, Government of Ontario, Ontario. See Mammals Section. Trichet, P., B. Boisaubert, H. Frochot, and J. F. Picard. 1987. Impact of herbicide treatments against bramble Rubus fruticosus Lagg on roe deer Capreolus capreolus L. Gibier Faune Sauvage 4: 165-88. See Mammals Section.

100.

101.

102.

33

Biodiversity, Conservation, and Habitat Restoration/Alteration 103. Truro, N. S. 1989. Controlling competing vegetation with lower than recommended rates of glyphosate. Forest research report Nova Scotia Forest Research Section . Habitat restoration/ glyphosate/ plant competition. 8 p. Vallentine, J. F. 1983. The application and use of herbicides for range plant control. Managing Intermountain Rangeland - Improvement of Range and Wildlife Habitats. S. B. Monsen, and N. Shaw (compilers). United States Department of Agriculture, Forest Service. General Technical Report INT-157. pp. 39-49. (*) Herbicides are an effective, necessary, and environmentally sound tool for the control of weeds and brush on rangelands to enhance livestock and big game forage availability. Various herbicide application methods and safety precautions are discussed. The properties and rangeland/pasture uses of 15 herbicides is presented in tabular form. Uses of amitrole, atrazine, dalapon, dicamba, 2,4-D, fenac, glyphosate, karbutilate, paraquat, picloram, silvex, 2,4,5-T, 2,3,6-TBA, tebuthiuron, and triclopyr in range management are discussed. VanSchip, W. M. 1990. Control of residual poplar and poplar re-growth on former jack pine dominated sites in the Brightsands F.M.A. Forest. Ministry of Natural Resources, Northwestern Ontario Forest Technology Development Unit , Thunder Bay, Ontario. Establishment report 4. 14 p. Harvesting and mechanical site preparation of northwestern Ontario jack pine sites may favor the increased regeneration of trembling aspen on those sites. Controlling the amount of aspen allowed to regenerate on former jack pine dominated sites is an important concern in current silvicultural strategy, but biological and time limitations of aerial tending with the presently registered herbicides (notably glyphosate) are a major challenge to the ability of foresters to manage aspen. This report describes the initiation of studies at the jack pine dominated Brightsands forest in the Thunder Bay district. The study is to evaluate six post-harvest manual herbicide ground application treatments for controlling aspen regrowth compared to aerial chemical site preparation with Vision (glyphosate). The report gives information on the experimental design, treatments to be used, plot location and design, and project schedule. Westbrooks, R. G. 1991. Plant Protection Issues. I. A commentary on new weeds in the United States. Weed Technology 5: 232-37. New or recently introduced weeds are biological pollutants in our natural and agricultural ecosystems. Unlike chemical pollutants, new weeds left unchecked often proliferate and pose problems that may not become apparent until eradication is too expensive or impractical. Management strategies for weeds should include: 1. prevention (from entering foreign commerce); 2. exclusion (detection of weed contaminants in imported products at ports of entry); 3. detection, containment, and eradication of incipient infestations; and 4. perpetual control (of widespread species that cannot otherwise be addressed). Appropriate legislative authority, modern weed technology, funding, and a renewed commitment to the concept of prevention are needed to prevent the introduction of new weeds. A national initiative to prevent the establishment of new weeds would be beneficial by saving on future losses and perpetual control costs. Actions taken now will prevent the continued introduction and spread of new weeds in the United States. Wiese, A. F., W. L. Harman, and C. Regier. 1994. Economic evaluation of conservation tillage systems for dryland and irrigated cotton (Gossypium hirsutum) in the southern Great Plains. Weed Science 42, no. 2 : 316-21. This 4-yr experiment was conducted to develop a profitable conservation tillage system for dryland and furrow-irrigated cotton in a winter wheat-fallow-cotton cropping system at Etter, TX. In the 2-yr cropping sequence, winter wheat was planted immediately after cotton harvest in mid-November and furrow irrigated. Three residual herbicide treatment combinations with atrazine or propazine with fluometuron, or propazine alone were sprayed on stubble after wheat harvest in late June and compared to repeated applications of glyphosate and conventional disk tillage. Glyphosate was used to control weeds that escaped residual herbicides. The following March, fluometuron, prometryn, and 2,4-D were sprayed on no-tillage treatments and trifluralin was incorporated with conventional planting. Hoeing cost ha-1 to control Palmer amaranth and kochia averaged about 13 where trifluralin was used, 34

104.

105.

106.

107.

Biodiversity, Conservation, and Habitat Restoration/Alteration 38 when initial treatment included fluometuron, 63 with propazine, and 93 when glyphosate alone was used to control weeds in fallow. Lint yield on dryland was about 390 kg/ha with disking and when glyphosate was followed by disking in spring, and 540 kg/ha with no-tillage. With irrigation, lint yield was 660 kg/ha with disking and 759 kg/ha with the best no-tillage treatment. Long-term profits/ha on dryland ranged from 340 for disking to 524 for propazine no-tillage. With irrigation, profit/ha ranged from 707 for disking alone to 751 when glyphosate was used after wheat harvest and followed by disking and incorporation of trifluralin the next spring before planting cotton. 108. William, S. I. 1995. Ten-year development of Douglas-fir and associated vegetation after different site preparation on Coast Range clearcuts. U.S.D.A. Forest Service Research Paper PNW 0 473. pp. 1-115. Wilson, S. D., and A. K. Gerry. 1995. Strategies for mixed-grass prairie restoration: Herbicide, tilling, and nitrogen manipulation. Restoration Ecology 3, no. 4 : 290-298. Large areas of North American prairie have been planted with grasses introduced from Eurasia. We examined three strategies (herbicide, tilling, and nitrogen manipulation) for enhancing the establishment of seedlings of native species and suppressing the introduced grasses Agropyron cristatum (crested wheat grass) and Bromus inermis (smooth brome). Plots (5 times 15 m) were subjected to one of three levels of tilling (non, intermediate, complete) and four levels of nitrogen (none, intermediate, high, and sawdust added to immobilize nitrogen). Treatments were applied in a factorial design with twelve treatments and ten replicates. Seeds of 41 native species were drilled into the plots in May 1992. Following the failure of seeds to establish in 1992, a subplot (5 times 13 m) within each main plot was sprayed with the herbicide glyphosate in April 1993. The nitrogen treatments were repeated in Spring 1993. In August 1993, the density of native seedlings in sprayed subplots was 20 times that in unsprayed subplots. Within sprayed subplots, native seedling density and the cover of bare ground decreased significantly with increasing nitrogen availability. Plots receiving sawdust had significantly higher mean cover of bare ground and significantly lower concentrations of soil available nitrogen. Native seedling density was significantly higher in plots receiving the highest intensity of tilling. The responses of native seedlings to all these factors point to the importance of neighbor-free establishment sites as a prerequisite for prairie restoration. Witt, J. S., A. S. Johnson, P. M. Dougherty, K. V. Miller, P. B. Bush, and J. Taylor. 1991. Effects of site preparation methods on vegetation structure assessed by two methods. Proceedings Southern Weed Science Society Perception: Fact or Fiction 44th Annual Meeting . Vegetation structure on various chemically and mechanically site-prepared pine plantations, 24 years post-treatment, was assessed with density board and light sensor techniques. Photographs of a density board were taken at 4 and 8 m and foliage densities at 4 height strata were determined. Light sensor measurements were recorded at 3 heights and percent light penetration was calculated as an index of vertical vegetation density. Differences among treatments occurred at 3 and 4 years posttreatment. Triclopyr and intensive mechanical treatments at 3 years post-treatment consistently had lower horizontal and vertical vegetation density than control or hexazinone treated plots. At 4 years post-treatment, differences were found in horizontal structure at <1.0 m and in vertical structure at >1.0 m. Differences in structural diversity among various site preparation methods were greatest at 3 years post-treatment. Wood, D. E., K. V. Miller, and D. L. Forster. 1996. Glyphosate and Fluridone for control of giant cutgrass (Zizaniopsis miliaceae) in waterfowl impoundments. Proceedings of the Annual Conference of Southeastern Association of Fish and Wildlife Agencies. 50: (In press). See Birds Section. Woodall, S. L. 1982. Herbicide tests for control of Brazilian-pepper and Melaleuca in Florida. Southeastern Forest Experiment Station, Asheville, North Carolina. USDA Forest Service Research Note. SE-314. 8 p. Container-grown Brazilian-pepper and melaleuca seedlings, 1-1/2 and 3-1/2 months old, were treated with eight herbicides at two concentrations in greenhouse screening trials. Root-absorbed herbicides were more effective than those foliar-absorbed, and gave virtually complete control with

109.

110.

111.

112.

35

Biodiversity, Conservation, and Habitat Restoration/Alteration both species. Herbicides with little residual activity were ineffective. Mature Brazilian-pepper bushes were field-tested with six herbicides and five application techniques. Basal spot treatments with Hyvar and Velpar gave good control. Stem injection of Tordon also gave good control, but was laborintensive. Stump treatments with silvex, Banvel, and Ammate reduced resprouting from stumps to acceptable levels, but did not adequately prevent root suckers. Foliar sprays of silvex, Banvel, Ammate, and Velpar gave erratic results, and generally were unsuccessful.

36

Birds

Birds
1. Batt, B. D. J., J. A. Black, and W. F. Cowan. 1980. The effects of glyphosate herbicide on chicken egg hatchability. Canadian Journal Zoology 58: 1940-1942. Hatchability and time to hatch of chicken eggs were found to be unaffected by application of glyphosate herbicide at three different concentrations and at four different embryo ages. We concluded that the use of the chemical as a weed control agent in zero tillage farming should not negatively affect the hatchability of the eggs of upland nesting birds. Bergvinson, D. J., and J. H. Borden. 1992. Enhanced woodpecker predation on the mountain pine beetle, Dendroctonus ponderosae Hopk., in glyphosate-treated lodgepole pines. The Canadian Entomologist 124 : 159-65. Lodgepole pines, Pinus contorta var. latifolia Engelm., treated with the herbicide glyphosate applied by axe frill or drill hole into the sapwood around the root collar, were readily infested by mountain pine beetles, Dendroctonus ponderosae Hopk. On trees treated with 360 mg of glyphosate per centimetre of circumference, foraging woodpeckers, Picoides spp., debarked 15% of the bole surface within 10 weeks and >30% after 1 year, compared with <5% for controls. Foraging efficiency on mountain pine beetle at 4 m exceeded 90% after applications by axe frill of glyphosate at doses of 360, 36, and 3.6 mg per centimetre of circumference, compared with <50% for controls. Glyphosatetreated trees rapidly became suitable for cavity excavation by woodpeckers. Blixt, D. C. 1993. "Effects of glyphosate-induced habitat alteration on birds, using wetlands." M.S. Thesis , North Dakota State University. Fargo. 127 p. The effects of wetland habitat alteration [50%, 70%, and 90% aerial coverage with glyphosate herbicide or untreated controls] on the populations of 8 bird species and 4 bird groups were studied in northeastern North Dakota, during June, July, and August 1990-1992. ... The herbicide successfully reduced live emergent vegetation in all treated wetlands. The 70% and 90% aerial spray coverages negatively influenced numbers of marsh wrens (Cistothorus palustris) and red-winged blackbirds (Agelaius phoeniceus). Numbers of common yellowthroats (Geothlypis trichas) decreased following 90% spray coverage. Generally, waterfowl (Anatidae) numbers did not differ among treatments. The numbers of yellow-headed blackbirds (Xanthocephalus xanthocephalus), song sparrows (Melospiza melodiai), mallards (Anas platyrhynchos), blue-winged teal (A. discors), gadwalls (A. strepera), rails (Rallidae), and sparrows (Emberizinae) did not respond to the herbicide treatments. Shorebirds (Scolopacidae, Charadriidae, and Recurvirostridae) generally showed higher abundances in the herbicide-treated wetlands than in the control wetlands during July and August. Thus, glyphosate may be a useful management tool for restoring cattail-choked wetlands. Cayford, J. T. 1988. The control of heather (Calluna vulgaris) with glyphosate in young forest plantations: the effect on habitat use by black grouse (Tetrao tetrix) in summer. Aspects of Applied Biology 16: 355-62. Habitat use by black grouse was studied in a 6-year-old Picea sitchensis plantation in North Wales, part of which had been sprayed with glyphosate in summer 1985. The effect of herbicide on the abundance of grouse food plants was assessed using quadrats. Glyphosate significantly reduced the amount of C. vulgaris and increased the amount of bare ground. Two years after spraying there was no difference in the abundance of Vaccinium myrtillus and Empetrum nigrum in treated and untreated areas. Telemetric observations and faecal samples both revealed a significant preference amongst adult black grouse in summer for areas sprayed with glyphosate. Habitat preference was not explained by the availability of food plants, but was probably related to their accessibility or fruiting quality. The results are discussed in relation to future weeding practices in conifer plantations, and the management of black grouse in commercial forests. (Wales, U.K.).

2.

3.

4.

37

Birds 5. Eggestad, M., E. Enge, O. Hjeljord, and V. Sahlgaard. 1988. Glyphosate application in forestecological aspects. VIII. The effect on Black Grouse (Tetrao tetrix) summer habitat. Scandinavian Journal of Forest Research 3: 129-35. Habitat preference of adult black grouse (T. tetrix) in relation to glyphosate spraying of conifer plantations was studied in SE Norway. Ten T. tetrix, six cocks and four hens, were fitted with radio transmitters and followed from late May through July 1985. Although we were not able to establish statistical significance, it appears that both cocks and hens avoid sprayed plantations the first one to two years after treatment, but after four to six years the plantations are used regularly by hens; the material on cocks is inconclusive. On certain sites vegetation development after spraying may be favourable to hens, but glyphosate spraying appears generally unfavourable to cocks. Therefore spraying plantations close to the leks is not recommended, as these are preferred summer habitats of the lek-cocks. (Norway). Environment Canada. 1990. Pesticide research and monitoring annual report 1988-1989. Environment Canada , ISBN 0-662-18461-0; Publication no: DSS cat no En 40-11-13-1989E. See Water Quality Section. Evans, D. D., and M. J. Batty. 1986. Effects of high dietary concentrations of glyphosate (Roundup) on a species of bird, marsupial and rodent indigenous to Australia. Environmental Toxicology and Chemistry 5: 399-401. Five zebra finches, Phoehila guttata, allowed unrestricted access to seed containing 5000 m.g glyphosate/g all died in 3-7 days, but they may well have died from starvation since their food consumption was drastically reduced. Six finches survived after ingesting seed containing 2500 m.g glyphosate/g for 5 days. The marsupial Sminthopsis macroura, and 2 species of hopping mouse, Notomys alexis and N. mitchelli survived on a diet in which the concentration of glyphosate was increased from 625 m.g/g to 5000 m.g/g by doubling the concentration of glyphosate in the food every few days during a 23-day period. The only toxic effect observed in the mammals was a marked body weight loss in the treated N. alexis. Thus, glyphosate for the 4 species is probably nontoxic to slightly toxic. (Australia). Flickinger, E. L., and G. W. Pendleton. 1994. Bird use of agricultural fields under reduced and conventional tillage in the Texas Panhandle. Wildlife Society Bulletin 22 : 34-42 . See Biodiversity and Habitat Restoration Section. Freedman, B. 1991. Controversy over the use of herbicides in forestry, with particular reference to glyphosate usage. Journal of Environmental Science and Health C8: 277-86. See Mammals Section. Hardy, B., and J. Desgranges. 1990. valuation des effets moyen terme sur les communauts aviennes de l'entretien des plantations d'pinettes noires (Picea mariana) aux phnoxys (Estaprop) et au glyphosate (Roundop). Serie de Rapports Techniques No. 101. Environment Canada, Canadian Wildlife Service. Effects over the medium term (seven to nine years) on avian communities of the maintenance of black spruce (Picea mariana) plantations with phenoxys (ESTAPROP) and glyphosate (ROUNDUP) were assessed during the breeding period in the Lower St. Lawrence and Gasp regions of Qubec. The study covered four vegetation parcels in which counts of breeding birds were conducted in the spring of 1988. Vegetation profiles for the parcels were completed over the summer of 1988. In the Lower St. Lawrence Region, the study focused on a black spruce (Picea mariana) plantation dating from 1980 and treated with glyphosate (Roundup) in 1981; a naturally regenerating parcel served as a control. In the Gasp, the selected plantation dated from 1976 and had been treated with phenoxys (Estaprop); an untreated plantation was available as a control. The bird count was conducted using fixed-diameter circular plot censuses (Blondel et al., 1973). Twenty plots were established in each plantation. Censuses were taken in each plot three times in the early morning and once at the end of the day in order to determine the population of all breeding bird species. The vegetation of each plot was profiled within one or several representative quadrats. 38

6.

7.

8.

9.

10.

Birds In the Lower St. Lawrence Region, 26 avian species were encountered in the treated plantation and 28 in the natural regeneration parcel. Nine (9) species exhibited significantly higher densities (p<0.05 or p<0.001) where glyphosate had been applied, as opposed to 1 species in the case of the naturally regenerating sites. Indices for avian diversity and dominance as well as the mean number of species and pairs per census plot were higher in the plantation. In the Gasp, 26 species were counted in the treated plantation and only 17 in the untreated plantation. Five (5) species had higher densities in the phenoxy-treated sites (p<0.05); however, six other species showed significant negative differences, being less abundant or absent within the treated plantation. The mean number of avian species per census plot was practically the same in both parcels, although the number of pairs and diversity index were higher in the treated plantation. On the other hand, the dominance index was higher in the untreated plantation. The vegetation parcels treated with phytocides exhibited more extensive coniferous cover than the control plots and more open spacing, while the untreated plantation was invaded by deciduous species and developed a dense foliage canopy. These changes to the habitat yield a significant rise in the density of several breeding bird species and an increase in avian diversity, primarily because of the mixed composition of the stands after seven to nine years of phytocide treatments. Over the medium term, the use of phytocides thus appears to be beneficial for the avian communities. 11. Henry, C. J., K. F. Higgins, and K. J. Buhl. 1994. Acute toxicity and hazard assessment of Rodeo, X-77 Spreader, and Chem-Trol to aquatic invertebrates. Archives of Environmental Contamination and Toxicology 27 , no. 3 : 392-99 . See Aquatic Invertebrates Section. Hoffman, D. J., and P. H. Albers. 1984. Evaluation of potential embryotoxicity and teratogenicity of 42 herbicides, insecticides, and petroleum contaminants to mallard eggs. Archives of Environmental Contamination and Toxicology 13: 15-27. The embryotoxicity of 42 environmental contaminants applied externally to mallard (Anas platyrhynchos) eggs including crude and refined petroleum , and complete formulations of herbicides and insecticides, are reported. Many of the petroleum pollutants were embryotoxic and moderately teratogenic and had LD50s of 0.3-5 m.L/egg (approximately 6-90 m.g/g egg). The most toxic was a commercial oil used for control of road dust followed by South Louisiana crude oil, Kuwait crude, no. 2 fuel oil, bunker C fuel oil, and industrial and automotive waste oil. Prudhoe Bay crude, unused crankcase oil, aviation kerosene, and aliphatic hydrocarbon mixtures were less toxic (LD50s of 18 to > 75 m.L) and less teratogenic. The median lethal concentrations (LC50s) of herbicides and insecticides in aqueous emulsion were measured by egg immersion: the most toxic were paraquat (1) and trifluralin (LC50s of approximately 1.5 lbs/acre; 1.7 kg/ha). Propanil, bromoxynil with MCPA, Methyl diclotop, Prometon, endrin, sulprofos, and parathion were toxic (LC50s of 7-40 lbs/acre; 7.8-44.8 kg/ha), whereas 2,4-D, glyphosate, atrazine, carbaryl, dalapon, dicamba, methomyl, and phosmet were only slightly toxic or not toxic (LC50s of 178 to > 500 lbs/acre; 199-560 kg/ha). Pesticides in nontoxic oil vehicle applied by microliter pipet were up to 18 times more toxic than when applied in water vehicle, which was probably due to better penetration of the pesticide past the eggshell and its membranes. Teratogenic effects and impaired embryonic growth are reported and results discussed in terms of potential hazard at field levels of application. A discussion is provided on the effects of pollutants on the eggs of other species of birds under laboratory and field conditions. Latka, R. J. 1992. An assessment of potential effects on least terns, piping plovers, and pallid sturgeon resulting from use of Rodeo and X-77 surfactant for tern and plover habitat enhancement. U.S. Army Corps of Engineers, Omaha District. 24 p. Interior least terns (Sterna antillarum) and piping plovers (Charadrius melodus) will not be adversely affected by fall or spring spraying of Rodeo. Fall spraying will be done after the birds migrate, and the chemical will have naturally biodegraded prior to their return in the spring. Spring spot-spraying can be accomplished prior to initiation of nesting. Early spraying will not affect the least terns or their food base of small fish. Small fish may inhabit shallow waters adjacent to the islands, but water will not be affected by spot-spraying. Spring spot-spraying prior to nesting may result in a small percentage of the insect food base utilized by piping plovers being inadvertently sprayed. Calculations of a "worst case" scenario (with excess Rodeo being sprayed over their entire food base) show that the 39

12.

13.

Birds chemical concentrations potentially ingested by plovers would not approach toxic concentrations for bird species. Spraying of Rodeo in the fall or in the spring will not adversely affect the pallid sturgeon (Scaphirhynchus albus). Spring spot-spraying will be done with hand-held equipment, so there will be little opportunity for chemical drift into adjacent waters. Most fall spraying will be broadcast using 4wheel all-terrain vehicles with boom attachments, reducing drift. Although aerial spraying will be avoided, it can be safely conducted. Even if some of the chemical would drift into the water, theoretical concentrations of glyphosate in aerially sprayed water would be extremely low, far below toxic concentrations for aquatic organisms. Rodeo was specifically designed for use in and around aquatic water bodies. Glyphosate biodegrades rapidly in water, especially flowing water, so waterborne Rodeo concentrations would be a temporary condition. Pallid sturgeon generally prefer deeper water, which would have greater dilution of the Rodeo than shallow areas adjacent to the islands. The environmental effects of fall and spring spraying of Rodeo are limited to the plant community, seen as the alteration of the terrestrial plant community, and the potential alteration of the aquatic emergent plant community if fall aerial spraying is utilized. Human disturbance during spring spraying would be minimal and temporary and should not affect least terns or piping plovers. 14. Lautenschlager, R. A. Effects of conifer release with herbicides on wildlife. (A review with an emphasis on Ontarios forests). Ontario Ministry of Natural Resources, Sault Ste. Marie, Ontario. VMAP Forest Research Information Paper No.111. 23 p. See Biodiversity and Habitat Restoration Section. . 1986. "In: Is Good Forestry Good Wildlife Management? J. Bissonette (editor). Forestry, herbicides, and wildlife. Miscellaneous Publication No. 689. Maine Agriculture Experimental Station, University of Maine, Orono. See Mammals Section. . 1991. Response of wildlife in northern ecosystems to conifer release with herbicides. Cooperative Forestry Research Unit , University of Maine, Orono, Maine. Information Report 26. 12 p. Total songbird densities are seldom reduced significantly by conifer release treatments. Species that prefer brushy deciduous cover, such as Wilson's warbler and the common yellowthroat, were reduced, whereas species that tend to avoid brushy cover, such as white-crowned and whitethroated sparrows increased. Conifer release treatments have reduced densities of red-backed voles and, occasionally, Townsend chipmunks and some shrew species all associated with brushy deciduous cover, but increased densities of Oregon voles, Trowbridge shrews, and deer mice, which commonly invade recently disturbed areas. Interestingly, only studies that used kill -- (snap and pit) trapping to study small mammal population responses have reported small mammal density reductions associated with conifer release treatments. For moose, conifer release treatments reduced browse availability for up to 3 growing seasons after treatment. Deer use remained unchanged or increased following these treatments during the first growing season after treatment. Both moose and deer likely benefit from conifer release treatments as succession proceeds, 2 to 8 years after treatment, and possibly for much longer. Although much effort has gone into the study of the effects of herbicides used for conifer release on wildlife in northern coniferous forests, those studies often lacked replication, pre-treatment information, and/or were conducted for one or at most two growing seasons after treatment. Therefore, valid generalizations about the effects of conifer release treatments on songbirds, small mammals, and deer are few and must be limited to two growing seasons or less after treatment. . 1993. Response of wildlife to forest herbicide applications in northern coniferous ecosystems. Canadian Journal of Forest Research 23: 2286-99. See Mammals Section.

15.

16.

17.

40

Birds 18. Linz, G. M., D. L. Bergman, and W. J. Bleier. 1992. Progress on managing cattail marshes with Rodeo herbicide to disperse roosting blackbirds. Proceedings of the 15th Vertebrate Pest Conference J.E. Borrecco & R.E. Marsh (editors). University of California, Davis. pp. 56-61. In August and September 1989 and 1990, we aerially sprayed 8 cattail (Typha spp.) marshes with Rodeo herbicide to begin evaluating its use for fragmenting dense cattail stands used by roosting blackbirds (Icterinae). Treated marshes were effectively eliminated as roost sites for blackbirds. After 2 years, cattail densities in 4 marshes treated with Rodeo at 5.8-7.0L/ha were 87% lower than pretreatment densities (P = 0.0001). In 1990, we treated 4 marshes with Rodeo at 4.7 L/ha. One year later, 6% of the cattails survived in the sprayed areas. Of 7 groups of indicator birds, only marsh wren (Cistothorus palustris) and rail (sora, Porzana carolina and Virginia rail, Rallus limicola) populations were adversely affected by cattail fragmentation. These preliminary results led to an increased research effort to develop marsh management techniques aimed at eliminating blackbird roosts. Linz, G. M., D. L. Bergman, D. C. Blixt, and W. J. Bleier. 1994. Response of black terns (Chlidonias niger) to glyphosate-induced habitat alterations on wetlands. Colonial Waterbirds 17 , no. 2 : 160-167 . The Black Tern (Chlidonias niger) is considered an endangered species in some states and is a category two species for listing as a federally threatened or endangered species. In the northern Great Plains, cattails (Typha spp.) have over-grown many wetlands, contributing to the decline in numbers of Black Terns. We aerially sprayed wetlands with glyphosate herbicide to assess the influence of habitat changes on Black Terns. In 1990 and 1991, two separate experiments were initiated in northeastern North Dakota whereby wetlands were randomly assigned one of three spray coverages with glyphosate. In one experiment, Black Terns used treated wetlands more than untreated wetlands (P lt 0.1). In the other experiment, densities of Black Terns were similar between untreated and treated wetlands (P gt 0.1). A stepwise multiple regression indicated a significant positive relationship between the number of Black Terns and hectares of water and dead emergent vegetation in the wetland (r-2 = 0.51). Black Terns use of a wetland may be positively influenced by floating mats of dead cattails, live emergent vegetation, and open water. Linz, G. M., D. L. Bergman, H. J. Homan, and W. J. Bleier. 1995. Effects of herbicide-induced habitat alterations on blackbird damage to sunflower. Crop Protection 14 , no. 8 : 625-29 . In August 1992, we treated cattail-dominated wetlands in four 23-km-2 blocks with aeriallyapplied glyphosate herbicide (5.3 l/ha). Four other blocks of wetlands were left untreated (reference). We assessed the effects of cattail (Typha spp.) reduction on roosting blackbird (Icterinae) numbers and sunflower damage within the blocks. Blackbird numbers did not differ between posttreatment years (P = 0.453) or between treated and reference wetlands (P = 0.469), averaging 6227 +- 4185 (SE) birds per block. Sunflower damage within blocks was similar between posttreatment years (P = 0.250) and did not vary (P = 0.460) between treatments (hivin x = 2.9 +- 1.2%). However, positive linear relationships were detected between blackbird numbers (y) and hectares of live cattails (x) ((y = 442.2 x)) (P = 0.006) and between blackbird numbers (x) and kilograms of sunflower lost per hectare per year (y) ((y = 0.003 x)) (P = 0.0001). Cattail reduction appears to discourage roosting blackbirds and, thus, may reduce sunflower damage in adjacent fields. Linz, G. M., D. C. Blixt, D. L. Bergman, and W. J. Bleier. 1996. Response of ducks to glyphosateinduced habitat alterations in wetlands. Wetlands 16 , no. 1 : 38-44 . The effects of glyphosate herbicide-induced changes in wetland emergent vegetation (largely cattails, Typha spp.) on densities of ducks (Anatinae) were assessed in northeastern North Dakota. In 1990 and 1991, 17 cattail-dominated wetlands were randomly assigned to 0% (reference wetlands), 50%, 70%, or 90% aerial spray coverage with glyphosate herbicide. Densities of green-winged teal (Anas crecca), blue winged teal (Anas discors), gadwalls (Anas strepera), and ruddy ducks (Oxyura jamaicensis) were similar among treatments during both post-treatment years (P gtoreq 0.1). One year post-treatment, mallard (Anas platyrhynchos) and northern pintail (Anas acuta) abundances did not differ among treatments (P gtoreq 0.1), whereas two years post-treatment their abundances were greater in the sprayed wetlands than in the reference wetlands (P ltoreq 0.1). Densities of northern shovelers (Anas clypeata) and redheads (Aythya americana) differed among treatments in both post41

19.

20.

21.

Birds treatment years, with the 50% sprayed wetlands harboring more ducks than did the other three treatments. Percent coverage and hectares of open water were positively correlated with numbers of diving ducks (Aythyini and Oxyurini) (P ltoreq 0.1). Dabbling duck (Anatini and Cairinini) numbers correlated positively with hectares of open water and dead vegetation, whereas their numbers were negatively correlated with percent coverage of live vegetation (P ltoreq 0.1). Results of this study suggest that numbers of ducks were positively influenced by creating a mosaic of open water, live vegetation, and dead vegetation with glyphosate herbicide. 22. . 1996. Responses of red-winged blackbirds, yellow-headed blackbirds and marsh wrens to glyphosate-induced alterations in cattail density. Journal of Field Ornithology 67 , no. 1 : 167-76 . The effects of herbicide-induced changes in wetland emergent vegetation on densities of territorial male Red-winged Blackbirds (Agelaius phoeniceus), Yellow-headed Blackbirds (Xanthocephalus xanthocephalus), and Marsh Wrens (Cistothorus palustris) were assessed in northeastern North Dakota. In 1990 and 1991, 23 cattail-dominated wetlands were randomly assigned to 0% (reference wetlands), 50%, 70%, or 90% aerial spray coverages with glyphosate herbicide. Two years post-treatment, densities of redwings were higher in the reference wetlands (hivin x = 1.59 1 0.24 (SE)/ha) than in the 90% treated wetlands (hivin x = 0.55 0.14/ha, P = 0.063). Yellowheads were more abundant in reference wetlands (hivin x = 3.80 0.83/ha) than in treated wetlands (hivin x = 2.05 0.40/ha, P = 0.061). Likewise, wrens were more abundant in reference wetlands (hivin x = 2.21 0.27/ha) than in treated wetlands (hivin x = 0.66 0.13/ha, P = 0.001). Percent coverage of live emergent vegetation (largely cattails, Typha spp.) was positively correlated with blackbird and wren numbers (P ltoreq 0.1). Results of this study suggest that numbers of these three wetland-dwelling species were limited by altering cattail density. Staggering vegetation management treatments on large wetland complexes may help diversify the stages of cattail regeneration and provide heterogeneous nesting and foraging habitat for these birds. MacKinnon, D. S., and B. Freedman. 1993. Effects of silvicultural use of the herbicide glyphosate on breeding birds of regenerating clearcuts in Nova Scotia, Canada. Journal of Applied Ecology 30: 395-406. The effects of habitat changes caused by the spraying of herbicides on the breeding birds of regenerating clearcut woodland were studied for one pre-spray and four post-spray years. Aerial application of glyphosate caused large decreases in the abundance of vegetation. There was a substantial recovery of some plant taxa by the end of the second post-spray growing season, especially of Rubus spp. and various herbaceous angiosperms, and there was further recovery to the end of the fourth post-spray year. Between the pre-spray and first post-spray years, the densities of most common breeding species decreased on all treatment plots, including the reference. In the second post-spray year, bird abundance on the reference plot increased to about the pre-spray condition, while on spray plots abundance stayed similar to the first post-spray year, and was smaller than the pre-spray population. The two most-abundant species, white-throated sparrow and common yellowthroat, decreased on all spray plots up to the second post-spray year, and then substantially recovered by the fourth post-spray year. Song sparrow and Lincoln's sparrow declined on the reference plot among the four study years, while on the spray plots these species were most abundant in the second and fouth post-spray years. As the reference plot continued its post-cutting successional development, it was colonized by black-and-white warbler, red-eyed vireo, ruby-throated hummingbird and palm warbler. This change in avifauna was inhibited on the spray plots, because the herbicide treatment set the vegetation back to an earlier successional state, and onto a different successional trajectory which was probably more conifer-dominated. In an analysis of the avifauna data by detrended correspondence analysis, all of the pre-spray plot-years ordinate together, indicating a similarity of species composition and abundance. The reference plot-years ordinate with chronologically increasing scores on axis 1, reflecting avifaunal changes during post-clear-cutting succession. Axis 2 reflects post-spraying changes in the avifauna of spray plots, lasting over a 2-year period, followed by a successional, post-herbiciding development of the avian community. Axis 2 is weaker than axis 1 (eigenvalues 0.151 and 0.065, respectively), so the

23.

42

Birds avifaunal changes associated with post-clearcutting succession are more prominent in the ordination than changes associated with glyphosate spraying. 24. Messersmith, C. G., Z. J. Woznica, R. G. Lym, and K. B. Thorsness. 1994. Evaluation of aquatic herbicides for managing blackbird roosting habitat. USDA Animal and Plant Health Inspection Service. North Dakota State University, Fargo, ND. Final Report. See Biodiversity, Conservation, and Habitat Restoration/Alteration Section. Millikin, R. L. 1996. Effects of vegetation management for brush on forest birds. Canadian Journal of Forest Research (Submitted). This report compares bird population characteristics and habitat use on 5 sites that had undergone vegetation management in a range of intensities. Bird species diversity, territory size, reproductive success and foraging behaviour were all apparently related to the amount of available brush. Milton, G. R., and J. Towers. 1990. Relationships of songbirds and small mammals to habitat features on plantation and natural regeneration sites. Canadian Institute of Forestry, St. Mary's River Forestry-Wildlife Project. Report No. 7. 57 p. See Mammals Section. Morrison M.L., and E. C. Meslow. 1984. Effects of the herbicide glyphosate on bird community structure in western Oregon USA. Forest Science 30: 95-106. Vegetative changes induced by the herbicide glyphosate and the resultant habitat use of birds nesting on 2 clearcuts in western Oregon were studied. About 23% of total plant cover was initially damaged by aerial application of glyphosate. Most measures of vegetation on the treated site decreased relative to the untreated site 1 year after glyphosate application. By 2 years postspray, vegetation on the treated site had recovered to near prespray status. No difference in density of the bird community was evident between treated and untreated sites during all years of study, although individual species densities were modified. Several bird species decreased their use of shrub cover and increased their use of deciduous trees 1 year after treatment. By 2 years postspray, many species returned to prespray use of most measured habitat components. Results indicated that application of glyphosate can modify the density and habitat use of birds. Rodgers, R. D. 1983. Reducing wildlife losses to tillage in fallow wheat fields. Wildlife Society Bulletin 11: 31-38. In the wheat-fallow system [used on about 33 million ha in 17 states and 3 Canadian Provinces], surface tillage for spring weed control destroys ... nests and flightless young in stubble [including those of ring-necked pheasants (Phasianus colchicus), sharp-tailed grouse (Pediocetes phastianellus), shorebirds (Scolopacidae, Charadriidae), and ducks (Anas spp.)], and kills or injures many incubating adults. Undercutters without mulch treaders saved ... nests, permitted survival of many flightless young, and caused no known injury to incubating adults during the spring weed control operation. Subsequent nest success was high. ... Fallow weed control with safe herbicides holds potential for providing undisturbed stubble for nesting. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. Wageningen, NL: European Weed Res. Soc. pp. 315-21 See Water Quality Section. Santillo, D. J., P. W. Brown , and D. M. Leslie, Jr. 1989. Response of songbirds to glyphosateinduced habitat changes on clearcuts. Journal of Wildlife Management 53: 64-71. We examined breeding bird populations and habitats on glyphosate (nitrogenphosphonomethylglycine) (Roundup, Monsanto, St. Louis, Mo.)-treated and untreated clearcuts in north-central Maine. Treatment of clearcuts with glyphosate herbicide reduced the complexity of vegetation through 3 years post-treatment compared to untreated clearcuts. Total numbers of birds, common yellowthroats (Geothlypis trichas), Lincoln's sparrows (Melospiza lincolnii), and alder 43

25.

26.

27.

28.

29.

30.

Birds flycatchers (Empidonax alnorum) were less abundant (P < 0.05) on treated clearcuts than on untreated clearcuts. Songbird densities were correlated with habitat complexity, especially hardwood regeneration, foliage height diversity (FHD), and vegetation height. Leaving untreated patches of vegetation and staggering herbicide treatments on large clearcuts will maintain bird populations similar to those on untreated clearcuts. 31. Solberg, K. L., and K. F. Higgins. 1993. Effects of glyphosate herbicide on cattails, invertebrates, and waterfowl in South Dakota wetlands. Wildlife Society Bulletin 21 : 299-307 . We studied the feasibility and effectiveness of using glyphosate herbicide to create openings in dense cattail stands in semipermanent wetlands and its possible effects on ducks and aquatic invertebrates in northeastern South Dakota. The 1986 and 1987 duck and invertebrate populations in wetlands treated with 2 application patterns of Rodeo herbicide in 1985 were compared with those in control wetlands with complete cattail coverage and in control wetlands with a natural interspersion of cattail and open water. Rodeo effectively controlled cattails, but treatment duration was related to water regime. Breeding ducks and over-water duck nest densities were greater on wetlands with chemically-created openings. Our findings and information in a review of the literature (Solberg 1989) suggest that Rodeo (glyphosate) herbicide is an acceptable management tool to create wetland openings that enhance waterfowl use and production. Aquatic invertebrate abundance was greater in untreated cattails in Rodeo-treated wetlands than in control wetlands and was less in treated and natural openings than in adjacent untreated emergent cattails. However, we were unable to determine if the differences in invertebrate abundance were caused by direct mortality from the glyphosate treatment or to movements of some invertebrates from treated to untreated areas within wetlands. Spidso, T. K., and N. Hovik. 1984. A pilot experiment of the effect of glyphosate on the growth and survival of Capercaillie tetrao-urogallus chicks. Symposium on Tetraonid Energetics, (42) 41. (Finland). Thompson, D. G., D. G. Pitt, P. T. Buscarini, B. Staznik, D. R. Thomas, and E. G. Kettela. 1994. Initial deposits and persistence of forest herbicide residues in sugar maple (Acer saccharum) foliage. Canadian Journal of Forest Research 24: 2251-62. See Plant and Soil Residues Section. Wood, D. E. 1996. "Evaluation of the herbicides glyphosate and fluridone for use in waterfowl management impoundments. " M.Sc. thesis , University of Georgia, Athens. 34 p. From 1993 through 1995, I conducted an operational evaluation of glyphosate and fluridone herbicides for the control of giant cutgrass (Zizaniopsis miliaceae) in waterfowl management impoundments in coastal Georgia, USA. Normal management activities, including prescribed fire and impoundment flooding, were allowed to evaluate herbicide effectiveness as a component of comprehensive management. Two treatment impoundments and one control impoundment were chosen. On October 14, 1993, impoundment 4A was treated with glyphosate herbicide at a rate of 4.66 L/ha. Impoundment 3A was treated with fluridone on April 5, 1994 at a rate of 18 kg/ha. Six 0.04 ha permanent plots were established in each impoundment. Rooted vegetation was recorded by presence/absence. Pretreatment sampling occurred in September 1993, with post-treatment sampling occurring in September 1994 and 1995. Both glyphosate and fluridone significantly reduced giant cutgrass frequency. However, glyphosate was more effective. Second year post-treatment frequencies of desirable species were either equivalent to or greater than pretreatment frequencies. Fluridone was evaluated to determine environmental persistence and ecological effects. Residues at 180 days post-treatment were greater than 50% of peak residue levels. Detectable residues were found up to 365 days post-treatment. Aquatic invertebrate populations were sampled to determine susceptibility to fluridone application. No consistent trend could be found that would indicate a treatment related depression of invertebrate populations.

32.

33.

34.

44

Birds 35. Wood, D. E., K. V. Miller, and D. L. Forster. 1996. Glyphosate and Fluridone for control of giant cutgrass (Zizaniopsis miliaceae) in waterfowl impoundments. Proceedings of the Annual Conference of Southeastern Association of Fish and Wildlife Agencies. 50: (In press). We conducted an operational scale trial of the herbicides Sonar (fluridone) and Rodeo (glyphosate) to evaluate control of giant cutgrass (Zizaniopsis miliaceae) and effects on waterfowl food plants in moist-soil managed impoundments of the Altamaha Waterfowl Management Area, Darien, Georgia. Sonar and Rodeo reduced giant cutgrass frequencies both post-treatment years, although greater reduction occurred in the Rodeo-treated impoundment. Panic grass (Panicum spp.) frequency within the Rodeo-treated impoundment decreased the first year post-treatment. First-year frequency of flat sedges (Cyperus spp.) decreased in all impoundments when compared to pretreatment frequency. However, second-year frequency did not differ from pretreatment for the Sonar and Rodeo-treated impoundments. First-year smartweed (Polygonum spp.) frequency was lower in the Sonar-treated impoundment; second-year frequency was higher in the Rodeo-treated impoundment. During the second year, wild millet (Echinocloa spp.) frequency was higher in the Rodeo-treated impoundment when compared to the control. Early fall application of Rodeo appears promising for the control of giant cutgrass and the enhancement of waterfowl food plants. Woodcock, J. M., R. A. Lautenschlager, F. W. Bell, and J. P. Ryder. 1997. Conifer release alternatives affect songbird populations in northwestern Ontario. Forestry Chronicle (In press). The effects of conifer-release alternatives on songbird densities (determined by territory mapping, mist netting, and banding) in four young (four- to seven-year-old) spruce plantations were examined. Twenty to 38 (block dependent) species bred on each treatment block. Post-treatment data revealed no major treatment-related changes in breeding bird species composition. The mean density of all birds decreased from 6.9 pairs/ha pre-treatment, to 6.3 pairs/ha in the first post-treatment growing season (P<0.05). Mean densities of the 11 most common species increased by 0.35/ha on the control plots during the first post-treatment growing season, but decreased on treated plots by 0.11/ha (brush saw), 1.6 /ha (Silvana Selective), 0.14/ha (Release) and 0.72 /ha (Vision). Following the coniferrelease treatments, Chestnut-sided Warbler (Dendroica pensylvanica) had lower (P<0.05) mean densities on the brush saw- and Silvana Selective-treated plots than on the control plots and fewer (P<0.05) female birds were captured in the first post-treatment year.

36.

45

Fish

Fish
1. Anton, F. A., E. Laborda, and M. DeAriz. 1994. Acute toxicity of the herbicide glyphosate to fish. Chemosphere 28 , no. 4 : 745-53 . The objective of this study was to evaluate the acute toxicity of three commercial products with the herbicide glyphosate as active ingredient on two species of freshwater fish: Carassius auratus L., and Oncorhynchua mykiss W. Static bioassays were used for toxicity testing. The experimental concentrations of the commercial herbicides were more high than the recommended doses to control weeds, showing a very low toxicity on both species of fish, with the experimental conditions used. Probably the current use of this herbicide would not be too hazardous to these fish species. (Spain) . Beyers, D. W. 1995. Acute toxicity of Rodeo registered herbicide to Rio Grande silvery minnow as estimated by surrogate species: Plains minnow and fathead minnow. Archives of Environmental Contamination and Toxicology 29 , no. 1 : 24-26 . Use of Rodeo registered herbicide (53.8% glyphosate, active ingredient) to control vegetation in or along irrigation canals in the Rio Grande basin in New Mexico has become a concern because of potential for populations of a federally endangered species, the Rio Grande silvery minnow (Hybognathus amarus), to be exposed. To investigate the potential for adverse effects, toxicity of Rodeo registered herbicide was estimated by conducting 96-h renewal-acute toxicity tests with a closely related species, plains minnow (Hybognathus placitus), and a standard laboratory animal, fathead minnow (Pimephales promelas). Rodeo registered herbicide had no effect on survival of plains minnow or fathead minnow at concentrations as high as 1,000 mg/L; thus, no-observed-acute-effect concentrations (NOAEC) for both species were 1,000 mg/L. Direct estimates of toxicity of Rodeo registered herbicide to Rio Grande silvery minnow were not possible because the species was not available for testing. However, the similarity of responses by plains minnow and fathead minnow suggests that 1,000 mg/L Rodeo registered herbicide is an acceptable estimate of the 96-h NOAEC for Rio Grande silvery minnow. Chapman, P. M. 1989. Salmonid toxicity studies with Roundup. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 063, pp. 257-62. Laboratory toxicity studies were conducted to assess acute lethal toxicity of ROUNDUP herbicide to three salomonid species (rainbow trout, chinook and coho salmon) and acute sublethal toxicity to coho salmon in seawater challenge tests. The results of this testing indicate that during the course of normal usage at the manufacturer's recommended use-rates, ROUNDUP poses no acute toxicity hazard to salmonids nor will it affect seawater adaptation of coho salmon. Environment Canada. 1990. Pesticide research and monitoring annual report 1988-1989. Environment Canada , ISBN 0-662-18461-0; Publication no: DSS cat no En 40-11-13-1989E. See Water Quality Section. Feei, S. 1987. Evaluating acute toxicity of pesticides to aquatic organisms carp, mosquitofish and daphnids. Plant Protection Bulletin 29: 385-96. Acute toxicity of pesticides was evaluated by standard methods in carp (Cyprinus carpio), mosquitofish (Gambusia affinis) and daphnids (Scapholeberis kingi). EC50 (median effective concentration) or LC50 (median lethal concentration) values varied according to species and test method, but the relative toxicity tendency of the pesticides was consistent for the 3 organisms. Pyrethroid insecticides were the most toxic to all 3 organisms. The herbicides 2,4-D, bentazone, glyphosate and paraquat, and all fungicides studied were less toxic than thiobencarb and butachlor and the pyrethroid insecticides. (Taiwan).

2.

3.

4.

5.

46

Fish 6. Folmar, L. C. 1976. Overt avoidance reaction of rainbow trout fry to 9 herbicides. Bulletin of Environmental Contamination and Toxicology 15: 509-14. (*) The results of these studies suggest that rainbow trout would not be killed by, and would not avoid, relatively low concentrations of dalapon, 2,4-D (DMA), glyphosate, Aquathol K, diquat, or TCA. However, trout present in acrolein-treated water probably would not avoid a lethal exposure. Trout strongly avoided low concentrations of copper sulfate and 0.1 mg/L xylene; although these chemicals are nontoxic at the levels tested, their presence in the water may influence the selection of habitat by fish. . 1978. Avoidance chamber responses of mayfly nymphs exposed to eight herbicides. Bulletin of Environmental Contamination and Toxicology 19: 312-18. See Aquatic Invertebrates and Algae Section. Folmar, L. C., H. O. Sanders, and A. M. Julin. 1979. Toxicity of the herbicide glyphosate and several of its formulations to fish and aquatic invertebrates. Archives of Environmental Toxicology and Contamination 8: 269-78. Studies were initiated to determine the acute toxicity of technical grade glyphosate (MON0573), the isopropylamine salt of glyphosate (MON0139), the formulated herbicide RoundupR (MON02139), and the RoundupR surfactant (MON0818) to four aquatic invertebrates and four fishes: daphnids (Daphnia magna), scuds (Gammarus pseudolimnaeus), midge larvae (Chironomous plumosus), mayfly nymphs (Ephemerella walkeri), Rainbow trout (Salmo gairdneri), fathead minnows (Pimephales promelas), channel catfish (Ictalurus punctatus), and bluegills (Lepomis macrochirus). Acute toxicities for Roundup ranged from 2.3 mg/L (96-h LC50, fathead minnow) to 43 mg/L (48-h EC50, mature scuds). Toxicities of the surfactant were similar to those of the Roundup formulation. Technical glyphosate was considerably less toxic than Roundup or the surfactant; for midge larvae, the 48-h EC50 was 55 mg/L and for rainbow trout, the 96-h LC50 was 140 mg/L. Roundup was more toxic to rainbow trout and bluegills at the higher test temperatures, and at pH 7.5 than at pH 6.5. Toxicity did not increase at pH 8.5 or 9.5. Eyed eggs were the least sensitive life stage, but toxicity increased markedly as the fish entered the sac fry and early swimup stages. No changes in fecundity or gonadosomatic index were observed in adult rainbow trout treated with the isopropylamine salt or Roundup up to 2.0 mg/L. The aging of Roundup test solutions for seven days did not reduce toxicity to midge larvae, rainbow trout or bluegills. In avoidance studies, rainbow trout did not avoid concentrations of the isopropylamine salt up to 10.0 mg/L; mayfly nymphs avoided 10.0 mg/L of Roundup, but not 1.0 mg/L. In a simulated field application, midge larvae avoided 2.0 mg/L of Roundup. Application of Roundup, at recommended rates, along ditchbank areas of irrigation canals should not adversely affect resident populations of fish or invertebrates. However, spring applications in lentic situations, where dissolved oxygen levels are low or temperatures are elevated, could be hazardous to young-of-the-year-fishes. Gardner, S. C., and C. E. Grue. 1996. Effects of Rodeo registered and Garlon registered 3A on nontarget wetland species in central Washington. Environmental Toxicology and Chemistry 15, no. 4: 441-51 . Purple loosestrife (Lythrum salicaria) is an invasive wetland perennial that became established in northeastern North America in the early 1800s. Despite its designation as a noxious weed, its distribution has continued to expand. Treatment with herbicides is the most widely used means of controlling purple loosestrife. This study examined the nontarget effects of two herbicides, Rodeo registered and Garlon registered 3A, currently used or being considered for use in controlling purple loosestrife in Washington State, respectively. Growth and/or survival of duckweed, Daphnia, and rainbow trout were monitored for at least 24 h following an application of each herbicide. Free-living water column and benthic invertebrates were monitored 24 h and 7 d post-spray using activity traps and sediment cores. Neither chemical was associated with significant decreases in survival or growth of the bioassay organisms, with the exception that growth of duckweed was reduced 48 h after exposure to Rodeo. Nor were significant decreases in the abundance of free-living aquatic invertebrates detected following the herbicide applications. Results suggest that neither herbicide, at the application rates used, poses a hazard to aquatic invertebrates in wetlands in central Washington.

7.

8.

9.

47

Fish However, Rodeo, because it is a broad-spectrum herbicide, may pose a greater hazard to nontarget aquatic vegetation. 10. Grande, M., S. Andersen, and D. Berge. 1994. Effects of pesticides on fish: Experimental and field studies. Norwegian Journal of Agricultural Sciences Supplement 13 : 195-209. (Norway) . Henry, C. J. 1992. "Effects of Rodeo herbicide on aquatic invertebrates and fathead minnows. " M.Sc. Thesis, South Dakota State University. See Aquatic Invertebrates and Algae Section. Hildebrand, L. D., D. S. Sullivan, and T. P. Sullivan. 1982. Experimental studies of Rainbow Trout populations exposed to field applications of RoundupR herbicide. Archives of Environmental Contamination and Toxicology 11: 93-98. This paper reports the effects of RoundupR herbicide (MON 02139) on rainbow trout viability and behaviour in several field experiments at the University of British Columbia Research Forest. Laboratory and field 96-hr LC50 values were similar: 54.8 and 52.0 mg/L. Avoidance-preference data indicated that fish would avoid lethal levels of Roundup. Operational application of Roundup at the recommended field dose (2.2 kg a.e./ha), as well as 10X and 100X field dose resulted in no mortality to rainbow trout in field streams. Results indicate that operational spraying with this herbicide for weed control should not be detrimental to rainbow trout populations. Improper use or accidental spills of Roundup could be avoided by rainbow trout and should not be lethal if diluted in a moderately-flowing stream. Holdway, D. A., and D. G. Dixon. 1988. Acute toxicity of permethrin or glyphosate pulse exposure to larval white sucker Catostomus commersoni and juvenile flagfish Jordanella floridae as modified by age and ration level. Environmental Toxicology and Chemistry 7: 63-68. A factorial design was used to determine the effects of age (2, 4, and 8 D for flagfish; 13, 20, and 26 D for white sucker) and ration (fed or unfed) on the acute toxicity of permethrin pulse exposure. A similar design was used to evaluate the acute toxicity of glyphosate to flagfish. Relative tolerance was assessed by determining the 2-H pulse-exposure concentration causing 50% mortality (PE LC50) over the subsequent 96 h. Age at exposure and the presence or absence of food modified the toxicity of permethrin to both flagfish and white sucker. White sucker were consistently less tolerant of permethrin than flagfish. Fed and unfed 8-day-old flagfish, as well as unfed, 2-day-olds, showed equivalent levels of permethrin tolerance, with respective 96-H PE LC50s of 0.57, 0.54, and 0.68 mg/l. They were significantly less tolerant than 4-day-old unfed juveniles (2.97 mg/l), which were in turn significantly less tolerant than 2- to 4-day-old fed juveniles, whose respective 96-H PE LC50s were 5.55 and 7.91 mg/l. Unfed white suckers were less tolerant than fed white suckers at all ages tested. Unfed white suckers at 13 days (96-H PE LC50, 0.002 mg/l) and 20 day (0.001 mg/l) were significantly less tolerant than unfed fish at 26 days (0.172 mg/l) and fed fish at 13 day (0.184 mg/l), which were more tolerant than 20-day-old fed fish (0.010 mg/l). Fed fish at 26 days of age were the most tolerant (3.668 mg/l). Glyphosate proved to be relatively non-toxic to flagfish up to 30 mg/l. No mortality was observed during the bioassays with 2- and 4-day-old fed and starved fish. Fed 8-day-olds (96-H PE LC20, 29.6 mg/l) were significantly more tolerant of glyphosate than were unfed 8-day-olds (96-H PE LC20, 2.94 mg/l). The sharply delineated periods of permethrin sensitivity are discussed with reference to the theory of saltatory ontogeny. Holtby, L. B. 1989. Changes in the temperature regime of a valley-bottom tributary of Carnation Creek, British Columbia, over-sprayed with the herbicide Roundup (R) (glyphosate). Forest Resources Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 063, 212-23. The temperature of a valley-bottom tributary of Carnation Creek was monitored for 2 years following the aerial application of the herbicide Roundup (glyphosate) to the riparian vegetation over the entire length of the tributary. Temperature changes were assessed by comparing water temperatures in the tributary with temperatures in the main-stem of Carnation Creek at a site immediately above the confluence of the main-stem and the tributary. Significant increases in mean

11.

12.

13.

14.

48

Fish daily temperature of 0.7 degrees C were observed during the summer following spraying. Daily maximum temperatures increased by 2.7 degrees C in the first summer after spraying and by 1.4 degrees C in the second summer. The increased temperatures observed after herbicide application were associated with the defoliation of riparian alder. Although no adverse effects of the temperature changes were observed on resident salmonids in the tributary, the additive effects of numerous herbicide treatments in the larger river could be of some concern, as would a similar treatment at a warmer site. 15. Holtby, L. B., and S. J. Baillie. 1989. Effects of the herbicide Roundup on Coho Salmon fingerlings in an over-sprayed tributary of Carnation Creek, British Columbia. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada , B.C. FRDA Report 063, pp. 273-85. The responses of coho salmon fingerlings to the over-spraying of a tributary of Carnation Creek with the herbicide ROUNDUP are described. Qualitative observations of caged coho fingerlings at exposed sites in the over-sprayed tributary indicated some stress 2 hours after application and within 24 hours of application some mortality (2.6%) of those fish was observed. Caged fish at other sites showed no similar signs of stress and no mortality immediately after application. We observed no unusual mortality of the approximately 120 coho fingerlings estimated to be resident in the tributary at the time of the herbicide application. In comparison with unaffected sites, catch per unit effort in the tributary declined immediately after the application but recovered within 3 weeks, suggesting that coho fingerlings had been stressed by some component of the herbicide spray. In comparison with 1-3 years of prespray data, in the two years following application no changes in overwinter mortality, growth rates, probabilities of entering and leaving the tributary or timing of spring emigration were observed that could be unambiguously attributed to the herbicide. Janz, D. M., A. P. Farrell, J. D. Morgan, and G. A. Vigers. 1991. Acute physiological stress responses of juvenile Coho salmon (Oncorhynchus kisutch) to sublethal concentrations of Garlon 4, Garlon 3A and Vision herbicides. Environmental Toxicology and Chemistry 10: 81-90. Juvenile coho salmon (Oncorhynchus kisutch) were exposed for 4 h to sublethal concentrations of the herbicides Garlon 4, Garlon 3A and Vision. Trials were performed in a closed-system respirometer that measured oxygen consumption of fish prior to and during a 4-h exposure. At the end of the exposure period, plasma glucose and lactate concentrations, hematocrit, and leucocrit were measured as indicators of acute physiological stress. There were no biologically significant indications of acute physiological stress in fish exposed to Garlon 4, Garlon 3A, or Vision at 5 to 80% of the 96-h LC50 concentrations. The results suggest that threshold herbicide concentrations causing physiological stress in short-term exposure (4 h) in juvenile coho salmon may be higher than the 96-h LC50 value for those herbicides. It was concluded that sublethal concentrations of Garlon and Vision herbicides do not induce significant physiological stress responses in juvenile coho salmon during a 4-h exposure period. Latka, R. J. 1992. An assessment of potential effects on least terns, piping plovers, and pallid sturgeon resulting from use of Rodeo and X-77 surfactant for tern and plover habitat enhancement. U.S. Army Corps of Engineers, Omaha District. 24 p. See Birds Section. Liong, P. C., W. P. Hamzah, and V. Murugan. 1988. Toxicity of some pesticides towards freshwater fishes. Malaysian Agricultural Journal 54: 147-56. The toxicity of 20 pesticides (Agrocide (lindane 65%), Desotox (DSMA 53.2%, diuron 4.5% with 2,4-D 11%), Dieldrex (dieldrin 15%), Dol G (lindane 6%), Furadan 2 G (carbofuran 2%), Gramoxone (paraquat 27.6%), Hopcin (fenobicarb 50%), Lebaycid 500 EC (fenthion 50%), Lindane 20 EC, Orthene 75 SP (acephate 75%), Roundup (glyphosate 30.5%), Sapterex (trichlorfon 80%), Sevin 85 WP (carbaryl 85%), Sogatox 22 (metolcarb 2% with phenthoate 2%), Sumidan 30 EC (fenitrothion with lindane), Thiodan (endosulfan 35%), Unden Dust (propoxur 1%), Unden 50 WP (propoxur 50%), Velpar L (hexazinone 25%), and Velpar 90 (hexazinone 90%)) to freshwater fish (Tilapia sp., Cyprinus carpio and Puntius javanicus) was examined in the laboratory. The LC50s for Tilapia after 24 h at 27-31oC were 75, 240, 0.13, 140, 24, 42, 8.3, 7.1, 0.56, no result, 7.4, 37, 14, 7.6, 1.1, 0.013, 1175, 32, 860 and 49

16.

17.

18.

Fish 500 ppm, respectively. (LC50 figures are also given for 48 and 96 h). The apparent no-effect levels were, for Tilapia, 5.6, 56, 0.032, 18, 5.6, 18, 1, 1.8, 0.18, 180, 1.0, 1.8, 1.8, 3.2, 0.32, 0.003, 180, 5.6, 320 and 320 ppm, respectively. The median lethal concentration ranged from 0.1 to 1000 ppm. Generally, the organochlorine insecticides were the most toxic. (Malaysia). 19. McLeay, D. 1988. Development of a bioassay protocol for evaluating the toxic risk to regional fisheries resources posed by forest-use herbicides. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 039, 57 p. Presently, no standardized approach exists for assessing the immediate or long-term toxic threat to anadromous and resident salmonid fish resources, posed by existing or candidate forestry herbicides. The types of aquatic toxicity tests currently being used are varied and provide no consistent means for comparing individual herbicides for relative toxicity. Nor do they necessarily relate to the freshwater life of primary concern (Pacific salmon and trout and their prey), or to the nature of exposure these species encounter when herbicides are applied within British Columbia forests. This report reviews the bioassay test procedures used to date for evaluating the toxicity of herbicides to salmonid fish and sensitive freshwater invertebrate species. Based on this review, and in consideration of regional needs as well as the most promising and applicable test procedures and approaches, a tentative bioassay protocol is developed and presented. Mitchell, D. G., P. M. Chapman, and T. J. Long. 1987. Acute toxicity of RoundupR and RodeoR to Rainbow Trout, Chinook and Coho Salmon. Bulletin of Environmental Contamination and Toxicology 39: 1028-35. (*) RoundupR, also known by the tradename VisionR, is a herbicide manufactured by Monsanto Company which may enter aquatic environments during the course of normal usage. This paper summarizes the results of aquatic toxicity testing of this herbicide using three different fish species, differentsized fish, and differences in dilution water (type/source, pH, hardness, and conductivity). There was no significant difference (p < 0.05) in the toxicity of Roundup/Vision to 0.37 g rainbow trout, 4.6 g chinook salmon and 11.8 g coho salmon. Acute toxicity values ranged from 7.4 mg/L to 12 mg/L in terms of the IPA salt of glyphosate, with 95% confidence limits ranging from 5.7-18 mg/L. Varying dilution water type (dechlorinated, dechlorinated and reconstituted, and natural lake water), pH (range 6.1-7.7), hardness (4.5-85 mg/L as CaCO3) and conductivity (12-132 mhos/cm) did not significantly affect Roundup/Vision toxicity to 0.37 g rainbow trout. The results of this study indicate that, according to a toxicity classification scheme currently in use, Roundup/Vision herbicide would be considered to be slightly toxic to trout and salmon species. Thus, no hazard to aquatic environments would be expected during the course of normal usage. . 1987. Seawater challenge testing of Coho Salmon smolts following exposure to Roundup herbicide. Environmental Toxicology and Chemistry 6: 875-78. Concentrations of Roundup herbicide of up to 10 times those encountered in the environment immediately after aerial application (2.78 mg/l) did not affect 24-h seawater survival; nor did plasma sodium concentration, haematocrit or growth in 10-day exposures with yearling Coho Salmon (Oncorhynchus kisutch). Similarly, no abnormal responses were noted when a 10-day freshwater recovery period was provided between the herbicide and seawater exposures. These results indicate that the application of Roundup at the manufacturer's recommended use-rates for forestry management will not adversely affect the ability of Coho Salmon smolts to adapt to seawater and to osmoregulate normally. Morgan, J. D., D. M. Janz, G. A. Vigers, and A. P. Farrell. 1989. Determination of threshold concentrations of forest-use herbicides causing rapid sublethal toxic effects towards salmonid fish. FRDA Project No. 2.46. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada , B.C.. A series of laboratory bioassay studies were conducted with juvenile rainbow trout and coho salmon to determine the acute lethal (96-h LC50) and behavioural and physiological threshold effect concentrations of the forest-use herbicides Garlon 3A, Garlon 4, and Vision with 10% and 15% surfactant formulations. Avoidance of the herbicides by rainbow trout did not occur until concentrations exceeded acute lethal 50

20.

21.

22.

Fish (96-h LC50) values determined for each herbicide. Nominal concentrations of the herbicides eliciting threshold avoidance reactions in test fish were 150 ppm Vision-10% surfactant, 54 ppm Vision-15% surfactant, 19.2 ppm Garlon 4 and 800 ppm Garlon 3A. These values represented respectively, 2.0, 2.0, 8.0 and 2.0 times the nominal 96-h LC50 values determined for each herbicide. Definite movements of test fish into herbicide solutions (i.e. preference reactions) were not observed. Overt behaviourial changes were observed in rainbow trout fry at levels 0.25 (Garlon 4) to 0.50 (Garlon 3A, Vision-10% and 15% surfactants) of the 96-h LC50 values. Nominal concentrations of the herbicides causing threshold changes in fish behaviour were 37.5 ppm Vision-10% surfactant, 13.5 ppm Vision15% surfactant, 0.60 ppm Garlon 4, and 200 ppm Garlon 3A. Thresholds for acute physiological stress responses in juvenile coho salmon were not reached for Vision (10% and 15% surfactants) and Garlon 4 up to the highest concentrations tested (60, 21.6 and 1.92 ppm, respectively). A threshold concentration of 200 ppm was determined for Garlon 3A, based on significantly elevated levels of plasma lactate. Based on significant differences in seawater survival, plasma Na+, haematocrit and condition factors between treatment and controls, nominal concentrations causing threshold effects in coho smolt seawater challenge tests were 7.5 ppm Vision-10% surfactant, 20.25 ppm Vision-15% surfactant, and 0.60 ppm Garlon 4. These values represented 0.10, 0.75 and 0.25 of the 96-h LC50 values determined for each herbicide, respectively. A threshold effect concentration could not be determined for Garlon 3A up to the highest concentration tested (200 ppm). 23. Morgan, J. D., G. A. Vigers, A. P. Farrell, and D. M. Janz. 1991. Acute avoidance reactions and behavioral responses of juvenile rainbow trout (Oncorhynchus mykiss) to Garlon 4R, Garlon 3AR and VisionR herbicides. Environmental Toxicology and Chemistry 10: 73-79. Laboratory bioassay experiments were conducted with juvenile rainbow trout (Oncorhynchus mykiss) to determine the acute lethal (96-h LC50) values of the forest-use herbicides Garlon 4R, Garlon 3AR and VisionR with 15% and 10% surfactant formulations, and the threshold concentrations of these herbicides that would cause behavioral effects. A Y-maze apparatus was used to determine the threshold concentration of each herbicide that elicited either an avoidance or a preference reaction from rainbow trout fry after a one-hour exposure period. Rainbow trout fry were observed for qualitative behavioural changes (e.g., increased coughing and ventilatory rates, loss of equilibrium, etc.) over a four-day period of exposure to a wide range of chemical strengths. Nominal concentrations of the herbicides eliciting threshold avoidance reactions in test fish were 150 ppm Vision-10% surfactant, 54 ppm Vision-15% surfactant, 19.2 ppm Garlon 4, and 800 ppm Garlon 3A, which were, respectively, 2, 2, 8, and 2 times the 96-h LC50 values determined for each herbicide. Preferential reactions to the herbicide solutions were not observed. Qualitative behavioral changes were observed in rainbow trout fry at levels from 25% (Garlon 4) to 50% (Garlon 3A, Vision-10% and -15% surfactants) of the 96-h LC50 values. Nominal concentrations of the herbicides causing threshold changes in fish behavior were 37.5 ppm Vision-10% surfactant, 13.5 ppm Vision-15% surfactant, 0.60 ppm Garlon 4, and 200 ppm Garlon 3A. Morgan, M. J., and J. W. Kiceniuk. 1992. Response of Rainbow trout to a two month exposure to R Vision , a glyphosate herbicide. Bulletin of Environmental Contamination and Toxicology 48: 772-80. (*) The 96hLC50 for Rainbow trout at 12 oC was 10.42 mg/L glyphosate in the Vision formulation. HPLC analyses confirmed the presence of glyphosate in the water, although the measured levels were lower than the nominal concentrations. This indicates that there was adsorption of the glyphosate onto the bucket walls and debris and/or uptake by the fish. Exposure of rainbow trout to glyphosate in the form of Vision at sublethal concentrations had little effect on the fish. There was no significant effect of exposure for either one or two months on any of the foraging variables that were measured. There was no significant effect of exposure to Vision for either one or two months on the time spent performing any agonistic activity. There was no significant effect on growth in length or weight of exposure to Vision for either one or two months. No livers, either from control or exposed fish, showed any evidence of tumors or melano-macrophages. All gill lesions were observed in small numbers, so the total number of lesions was used in the analyses. There was no significant difference between the number of lesions on gills of fish exposed to Vision and the control fish. There were significant differences between control and exposed fish in one aspect of agonistic behaviour. After one month of exposure fish in the highest concentration (45.75 g/L) had a

24.

51

Fish significantly higher frequency of wigwags. After two months, fish exposed to the lowest concentration of Vision performed significantly fewer wigwags. It is not known what the implications of a change in one agonistic activity in the repetoire of aggressive behaviour would be in terms of a fish's ability to hold a feeding territory. Also, the dose response was inconsistent over time, making interpretation of the results more difficult. It should be noted however, that this effect occurred at a concentration well below the level that has been measured following some spray operations. 25. Neskovic, N. K., V. Poleksic, I. Elezovic, V. Karan, and M. Budimir. 1996. Biochemical and histopathological effects of glyphosate on carp, Cyprinus carpio L. Bulletin of Environmental Contamination and Toxicology 56 , no. 2 : 295-302 . Glyphosate, also known by the trade names Roundup and Rodeo for agricultural use, is a broad-spectrum, translocated herbicide, used primarily in agricultural applications, and for vegetation control in non-crop areas. It is used as non-selective herbicide and for aquatic weed control in fishponds, lakes, canals, slow running water, etc. (USDA 1984). Glyphosate is perhaps the most important herbicide ever developed. Literature of toxicological and ectoxicological properties of glyphosate is extremely sparse, considering its importance as herbicide. Generally, glyphosate is slightly toxic to mammals and fish, but it may have an impact on the aquatic environment and also on the other aquatic organisms (USDA 1984). Due to this, its toxicity investigation is very important. The study of sublethal effects is of special importance for toxicological evaluation of compound. The objective of this study was to investigate acute and subacute toxic effects of sublethal glyphosate concentrations in water to carp (Cyprinus carpio L.), one of the commercially most important fish species population freshwaters of Yugoslavia. (Yugoslavia) . Newton, M., K. M. Howard, B. R. Kelpsas, R. Danhaus, C. M. Lottman, and S. Dubelman. 1984. Fate of glyphosate in an Oregon USA forest ecosystem. Journal of Agricultural and Food Chemistry 32: 1144-51. See Mammals Section. Olaleye, V. F., and O. A. Akinyemiju. 1996. Effects of glyphosate (N-(phosphonomethyl) glycine) application to control Eichhornia crassipes Mart. on fish composition and abundance in Abiala Creek, Niger Delta, Nigeria. Journal of Environmental Management 47 , no. 2 : 115-22 . Roundup, a commercial preparation containing 360 g/l glyphosate in the form of 480 g/l isopropylamine salt was applied to control water hyacinth (Eichhornia crassipes Mart.) in Abiala creek, Delta State, Nigeria. The herbicide, which was applied at a rate of 2.88 kg a.e./ha on a heavily infested stretch of the creek, succeeded in killing mats of water hyacinth in the treated plots within 7 days after treatment. The presence of floating aquatic vegetation caused a low fish catch in the untreated control plots during the period of study. The population of fish specimens caught in the glyphosate-treated plots increased significantly after treatment. The presence of juveniles and immature adult fishes in the treated areas of the creek indicated that the impaired water quality had been adequately regenerated 28 days after treatment. (Nigeria) Pawlizki, K. H. 1987. Effects of herbicides used in orchards and vineyards on man and the environment. Gesude Pflanzen 39: 486-96. See Plant and Soil Residues Section. Payne, N. J., J. C. Feng, and P. E. Reynolds. 1990. Off-target deposits and buffer zones required around water for aerial glyphosate applications. Pesticide Science 30: 183-98. See Aquatic Invertebrates and Algae Section. Payne, N., J. Feng, and P. Reynolds. 1989. Off-target deposit measurements and buffer zones required around water for various aerial applications of glyphosate. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 88-109. An investigation was conducted to quantify off-target deposit from three types of aerial glyphosate use-strategies for forestry use, and to estimate the width of buffer zones required around water to protect fish and their food supply from direct toxicological effects. To overcome the difficulty 52

26.

27.

28.

29.

30.

Fish of estimating different buffer widths to meet the various conditions encountered. e.g., windspeed, boundary layer stability, active ingredient application rate, etc., a realistic worst case scenario was chosen, and data were collected accordingly. A buffer width of 25 m around water bodies is adequate to protect salmon, rainbow trout and aquatic invertebrates from significant direct effects resulting from off-target Roundup deposits from the Microfoil and Thru Valve Boom use-strategies; for the D8-46 usestrategy, a 30 m buffer width is required. 31. Reynolds, P. E., J. C. Scrivener, L. B. Holtby, and P. D. Kingsbury. 1989. An overview of Carnation Creek herbicide study: historical perspective, experimental protocols, and spray operations. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, FRDA Report 063, pp. 1526. See Water Quality Section. . 1989. A summary of Carnation Creek herbicide study results. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA report 063, pp. 322-34. The Carnation Creek watershed, located on the west coast of Vancouver Island, was aerially treated with Roundup (glyphosate) from September 6 to 15, 1984 with the intent of examining environmental fate and impact of Roundup treatment on a temperate coastal rain forest. Carnation Creek presents a unique research situation in that a 20 yr data base on its salmonid population exists. This data base has permitted whole life history impact assessment of the effects of numerous forest management practices, including the possible impacts of herbicide use on this resource. Following Roundup treatment at Carnation Creek, various chemical and biological studies were conducted for up to three years post-treatment. These studies revealed no unexpected or long-term adverse effects on coho salmon or other aquatic organisms using tributaries that had been directly over-sprayed with the herbicide. Residue movements within the watershed and residue inputs into the aquatic ecosystem were carefully monitored in relation to autumn and winter storms. Glyphosate residues rapidly dissipated and degraded in the natural environment. After one year remaining residues were strongly absorbed to organic matter, soil particles, and/or stream bottom sediments, were deemed to be biologically unavailable. . 1993. Review and synthesis of Carnation Creek Herbicide research. The Forestry Chronicle 69: 323-30. See Water Quality Section. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. Wageningen, NL: European Weed Res. Soc. pp. 315-21. See Water Quality Section. Samis, S. C., S. von Schuckmann, M. T. Wan, G. D. McKellar, and M. Scott. 1992. Guidelines for the protection of fish and fish habitat during use of glyphosate and other selected forestry herbicides in coastal British Columbia. Canadian Manuscript Report Fish and Aquatic Science. No. 2176. 14 p. Chemical vegetation control in the forest industry in British Columbia is conducted primarily by means of aerial and ground application of the herbicide glyphosate and hand application of 2,4-D in selective individual tree treatments. These guidelines provide a standardized fish and fish habitat protection strategy for use by herbicide applicators and regulatory agencies responsible for fisheries and aquatic environmental protection. The strategy is based on an established system of stream classification and the provision of a ten metre pesticide free zone around watercourses where it is deemed necessary.

32.

33.

34.

35.

53

Fish 36. Servize, J. A., R. W. Gordon, and D. W. Martens. 1987. Acute toxicity of Garlon 4 and Roundup herbicides to Salmon, Daphnia, and Trout. Bulletin of Environmental Contamination and Toxicology 39: 15-22. (*) This report summarizes the acute lethality of Garlon 4, Roundup and a surfactant (MON0818) contained in the formulation of the latter to sockeye salmon (Oncorhynchus nerka), Daphnia pulex, rainbow trout (Salmo gairdneri) and coho salmon (Oncorhynchus kisutch). The mean 96-hour LC50 of Garlon 4 for sockeye was 1.3 mg/L at 4.5oC and pH 7.9, while the 96-hour EC50 for Daphnia was 1.2 mg/L at 21o C. The 96-hour LC50's for coho and rainbow trout were 2.2 mg/L at 15oC and pH 6.2. Sockeye fingerlings and fry were equally susceptible to Roundup; the 96-hour LC50 averaged 27.7 mg/L when tested at pH 7.7 to 8.0. For Daphnia the 96-h EC50 was 25.5 mg/L. The average 96-hour LC50 for rainbow trout was 26.8 mg/L; and for a single test of coho the 96-hour LC50 was 42 mg/L. The sum of toxic units for glyphosate and MON0818 ranged from 1.39 to 1.92 for the sockeye, Daphnia, rainbow and coho tested in moderately or weakly buffered waters. Singh, S. P., and N. K. Yadav. 1978. Toxicity of some herbicides to major carp fingerlings. Indian Journal of Ecology 5: 141-47. As a consequence of herbicide treatment in or near fish habitats, fish are likely to be killed due to direct effects of the herbicides and indirectly due to asphyxiation following the disappearance of vegetation. Ten herbicides were, therefore, evaluated for their acute toxicity to major carp (Cirrhina mrigala) fingerlings at concentrations of 1 to 100 ppm. Copper sulphate proved to be most toxic, caused 40 percent fish mortality at 1 ppm, and 2,4-D amine was least toxic, causing no mortality even at 100 ppm concentration. Other herbicides in the order of decreasing toxicity to fish were: paraquat > glyphosate > terbutryne > diuron > simazine > dalapon > diquat > MSMA. In the presence of coontail (Ceratophyllum demersum), a submerged aquatic plant, the recommended field rates of simazine, paraquat and diuron were found to be toxic to fish exposed for 24 days. (India). Tooby, T. E. 1981. Predicting the direct toxic effects of aquatic herbicides to nontarget organisms. Proceedings of a Symposium on Aquatic Weeds and their Control. Christ Church, Oxford. pp. 265-74. . 1985. Fate and biological consequences of glyphosate in the aquatic environment. In: The Herbicide Glyphosate. E. Grossbard, and D. Atkinson, 206-17. London: Butterworths. See Water Quality Section. Tooby, T. E., J. Lucey, and B. Stott. 1980. The tolerance of grass carp, Ctenopharyngodon idella Val., to aquatic herbicides. Journal of Fish Biology 16: 591-97. Because grass carp may feed selectively on water plants, additional conventional methods may be necessary for effective control of aquatic vegetation. For economic reasons aquatic herbicides are most likely to be used in conjunction with grass carp within an integrated aquatic weed control programme. The acutely lethal toxicities of ten herbicides (asulam, cyanatryn, 2,4-D amine, dalapon, dichlobenil, diquat, diuron, glyphosate, paraquat, and terbutryne) to grass carp were measured. A preliminary evaluation of the risk to these fish showed that, when used at the recommended rate, the maximum herbicide concentration likely to be found in water should cause little harm to the fish, although attention should be given to the possible inhibition of feeding. Trotter, D. M., M. P. Wong, and R. A. Kent. 1990. Canadian water quality guidelines for glyphosate., Water Quality Branch, Inland Waters Directorate, Environment Canada, Ottawa, Ontario. 27 p. See Water Quality Section. U.S.D.I. 1981. Fisheries and Wildlife Research 1980. U.S. Department of the Interior, Fish and Wildlife Service. 201 p. The acute toxicity of three widely used herbicides to fish and invertebrates was determined: atrazine at 24 ppm was not toxic to rainbow trout, bluegills or daphnids; neither was metribuzin at 40 ppm to rainbow trout and daphnids; nor purified 2,4,5-T at 100 ppm to daphnids, scuds or midges. In one experiment in South Dakota to investigate the effects of herbicides on aquatic plants and 54

37.

38.

39.

40.

41.

42.

Fish invertebrates in the northern prairie wetlands and pothole lakes, the presence of 50 ppm suspended sediment reduced the toxicity of glyphosate to daphnids to about onethird. Further studies indicated that 28% of the radioactivity accumulated by daphnids exposed to C14-PCP at 50 ppb was detected as polar, watersoluble metabolites. Dechlorination and methylation were the major pathways of PCP disappearance in hydrosoil interfaces. High vegetative biomass and photosynthetic rates mitigated the toxic effects of PCP, allowing increased survival and fecundity in several fish species. 43. Wan, M. T., R. G. Watts, and D. J. Moul. 1989. Effects of different dilution water types on the acute toxicity to juvenile Pacific Salmonids and Rainbow Trout of glyphosate and its formulated products. Bulletin of Environmental Contamination and Toxicology 43: 378-85. This study indicates that a notable variation of 96-h LC50 values is attained for the same fish species when different water types are used in the bioassay of glyphosate, MON0818, MON8709, and RoundupR. Variation of 96-h LC50 values for MON0818, MON8909, and RoundupR is in the same order of magnitude, irrespective of water types. For glyphosate, the 96-h LC50 values for different water types can vary by an order of magnitude. Hardness of water and pH appear to be the key factors causing the variation of 96-h LC50 values. RoundupR, MON8709, and MON0818 are more toxic to young salmonids in hard than they are in soft water, while the reverse holds for glyphosate. Salmonid 96-h LC50 data for RoundupR and MON0818, but not for glyphosate and MON8709, obtained from reconstituted water are generally similar to 96-h LC50 values generated from natural sources of dilution water. . 1991. Acute toxicity to juvenile Pacific northwest salmonids of Basacid Blue NB755 and its mixture with formulated products of 2,4-D, glyphosate, and triclopyr. Bulletin of Environmental Contamination and Toxicology 47: 471-78. (*) Basacid Blue NB755 (BB) is a triphenylmethane dye used in British Columbia mainly as a marker by aerial and ground herbicide spray operators in forestry. This bioassay study indicates that basacid blue dye is quite toxic to salmonids, particularly in soft water. This dye (with no herbicides) is not likely to have an acute impact on salmonids at rates commonly used by foresters. However, the use of this dye as an indicator at the rate of 100 mg/L in spray mixtures increases the toxicity to young salmon of formulated products of 2,4-D amine, 2,4-D ester, Garlon 4 (except in intermediate water) but not Roundup, irrespective of water types. It is suggested that the use of basacid blue dye indicator for now for Roundup ground and aerial operations should not exceed 100 mg/L (equivalent to 8 mL Basacid Blue NB755 per 100 L spray mixture) per spray mixture load of 100 L. Wan, M., R. Watts, and D. Moul. 1988. Toxicity to salmonids of the herbicide glyphosate (Roundup) and its surfactant. Regional Program Report 86-16. Department of the Environment. Environmental Protection, Pacific Region . In 1984, static fish bioassays were conducted by the Environmental Protection Service, Pacific Region and Yukon, of Environment Canada to determine the acute toxicity to juvenile salmonids of TM TM herbicide Roundup and its surfactant MON0818. This study indicated that Roundup was only moderately toxic to two species of Pacific salmon. The average 96 hr LC50 values for the fish were: Coho salmon (42.3 mg/L), and Rainbow trout (32.4 mg/L). However, the surfactant MON0818 was highly toxic to salmonids, with 96 LC50 values of 3.2 and 3.5 mg/L for Rainbow trout and Coho salmon, respectively. The stabilization time of LC50 values varied with different fish for RoundupTM, but appeared to be 96 hours for the surfactant MON0818. Analytical results of selected test water samples indicated that the free acid equivalent of glyphosate in RoundupTM was approximately 30%. Also glyphosate did not degrade to any significant degree during the 96-hour test period. The implications of these results and their potential short term impact on Pacific salmonid species are discussed. Wang, Y., C. Jaw, and Y. Chen. 1994. Accumulation of 2,4-D and glyphosate in fish and water hyacinth. Water, Air and Soil Pollution 74 , no. 3-4 : 397-403 . At a concentration of approximately one hundredth and one thousandth of the 48-hr LC sub(50) values, the accumulation of radioactivity and relative concentration of 2,4-D and glyphosate in carp and tilapia were studied by using labelled and unlabelled chemicals. About 83 (at a concentration of 0.5 ppm) and 91% (of 0.05 ppm) of the radioactive matter remained in the water until 14 days after super (14)C-2,4-D amended, but only 17.2% of glyphosate remained in the water with 0.05 ppm concentration 55

44.

45.

46.

Fish of glyphosate. No significant variation was shown in the accumulation of the concentration of herbicide in fish from 2 to 7 d. Although glyphosate disappeared within 3 d in water under sunlight, the radiochemicals in the water hyacinth remained constant to the 14th d. (Taiwan) .

56

Human Health

Human Health
1. Brewster, D. W., J. J. Warren, and W. E. Hopkins II. 1991. Metabolism of glyphosate in SpragueDawley rats: Tissue distribution, identification, and quantitation of glyphosate-derived materials following a single oral dose. Fundamental and Applied Toxicology 17: 43-51. See Mammals Section. Bronstein, A. C., and J. B. Sullivan. 1992. Herbicides, fungicides, biocides, and pyrethrins. Hazardous Materials Toxicology, Clinical Principles of Environmental Health. J. B. Sullivan, Jr. , and G. R. Krieger (editors). Baltimore, Maryland : Williams and Wilkins. pp. 1063-77. The toxicology and use of herbicides, fungicides, biocides, and pyrethrins were discussed. Topics included sources, production, exposures, clinical toxicology, diagnosis and treatment of poisoning, and exposure limits for the following: chlorophenoxy herbicides, paraquat (2074502), glyphosate (1071836), pentachlorophenol (87865), 2,3-dichloro-1,4-naphthoquinone (117806), benomyl (17804352), thiabendazol (148798), thiophanate-methyl (23564069), ferbam (14484641), thiram (137268), ziram (137304), acrolein (107028), organotin compounds, pyrethrins, and pyrethroids. The acute toxic effects of exposure to chlorophenoxy herbicides have induced gastroenteritis, skeletal muscle myotonia, myoglobinuria, cardiac dysrhythmias, and central nervous system depression. Chronic effects have included chloracne, increased incidence of certain cancers, and immune dysfunction. Ingestion of large amounts of paraquat (usually suicide) may cause death from cardiovascular collapse within a few days or from irreversible pulmonary fibrosis within weeks. Exposure to the fungicide pentachlorophenol has resulted in eye, nose, throat, mucous membrane, and respiratory irritation after inhalation and gastrointestinal symptoms after ingestion. Aplastic anemia has also been associated with exposure to pentachlorophenol. Canadian Centre for Occupational Health and Safety. 1990. Glyphosate. Hamilton, Ontario. Chemical safety information sheet. Herbicide. Toxicity: low oral toxicity: irritation of respiratory tract, skin and eyes. Centre de Toxicologie du Qubec . 1988. tude de lexposition professionnelle des travaileurs exposs au glyphosate. Ministre de Energie et des Ressources, Centre Hospitalier de Universit Laval. ER89-1110. 27 p. The present study was undertaken in order to evaluate the exposure of forestry workers to the herbicide glyphosate and to determine if the recommended safety practices were effective in limiting this exposure. Toxicity of glyphosate, whether acute or chronic, is low, as determined by mammalian studies. Skin, eye, and mucous membrane irritation is the only demonstrated problem. This project was carried out in three steps: the development of an analytical method for the determination of glyphosate in urine, its application to the biological monitoring of a small group of workers, and finally the monitoring of a larger number of workers in order to estimate the effectiveness of the safety practices in use. The analytical method developed uses high performance liquid chromatography (HPLC) with fluorescence detection. The detection limit of 15 g/L is adequate for the identification of overexposure to glyphosate. During the summer of 1986 we monitored a crew of forest workers (foreman, mixer, operator and two flagmen) in charge of glyphosate spraying operations. At the same time, the Ministry of Energy and Resources measured the levels of glyphosate in the breathing zone of the workers, using personal sampling pumps. Both methods showed the mixer to be the most exposed worker, however biological monitoring gave a more realistic assessment taking into account all routes of exposure. Medical examinations conducted on the workers showed no abnormalities. During the summer of 1987, a larger-scale study was conducted on 40 forest workers, mostly mixers, based in several regions of the province. Some of these worked for the MER, and others for private contractors. Each participant was asked to submit morning urine samples on six consecutive days during the spraying period, for the determination of glyphosate. All workers were then interviewed individually to evaluate the availability and use of safety equipment, and their attitude towards safety. 57

2.

3.

4.

Human Health Statistical analyses performed upon the data showed that the levels of glyphosate in the urine of government employees were significantly lower than those of workers hired by private contractors. The latter were not as well informed of the hazards involved in pesticide handling, made less use of safety equipment and generally lacked appropriate washing facilities. It was thus demonstrated that good safety practices are effective in reducing worker exposure to glyphosate. 5. Chan, P. C., and J. F. Mahler. 1992. NTP technical report on toxicity studies of glyphosate administered in dosed feed to F344/N rats and B6C3F1 mice. Toxicity Report Series No. 16. 55 p. NIH Publication No. 92-3135. The toxicity and disposition of glyphosate (1071836) was studied in rats and mice. Male F344/N-rats were gavaged with 5.6 or 56 mg/kg radiolabeled glyphosate. Urine and feces were collected at 24 hour intervals for 72 hours and analyzed for activity. Selected rats were killed 3 to 96 hours post dosing to determine the tissue distribution of radioactivity. Approximately 20 to 30% of either dose was eliminated in the urine and 70 to 80% in the feces over 72 hours. Only about 1% of the dose remained in the tissues, mostly in the liver and small intestine. Ten F344/N-rats and B6C3F1mice were administered 3125, 6250, 12,500, 25,000, or 50,000 parts per million (ppm) glyphosate in their diet for 13 weeks. Surviving animals were killed at the end of the study and necropsied. Blood samples were collected from the rats at necropsy to determine hematological and serum chemistry parameters. Glyphosate mutagenicity was evaluated in the Ames/Salmonella assay with or without S9 metabolic activation and by the mouse peripheral blood micronucelus test. All rats survived until the end of the study. Glyphosate doses of 12,500 ppm or higher caused slight increases in hematocrit and hemoglobin. Serum bile acid concentrations and alkaline-phosphatase and alanine-aminotransferase activities were significantly increased. One mouse treated with 50,000 ppm died. The major pathological change induced by glyphosate was a dose related increase in basophilic changes and hypertrophy of acinar cells (cytoplasmic alterations) in the parotid and submandibular salivary glands of rats and the parotid salivary glands in mice. All doses caused these changes in rats. In mice, doses of 6250 ppm or higher caused these effects. No histopathological changes were seen in the liver. Glyphosate was not mutagenic. Couture, G., J. Legris, and R. Langevin. 1995. valuation des impacts du glyphosate utilis dans le milieu forestier, Qubec. Gouvernement du Qubec, ministre des Ressources naturelles, Direction de lenvironment forestier, Service du suivi environmental , Qubec. RN953082. 187 p. See Plant and Soil Residues Section. Dawson, M. A., and J. L. Renfro. 1993. Effects of pesticides on organic anion secretion by winter flounder renal proximal tubule cells in culture. 6th International Symposium on Responses of Marine Organisms to Pollutants. J. J. Stegeman, M. N. Moore, and M. E. Hahn (editors). Woods Hole, MA. Marine Environmental Research 35:224-25 A number of toxic compounds, both endogenous and xenobiotic, are secreted as organic anions in the proximal tubule. Such secretion helps determine total body content of toxic substances and may also subject the kidney to damage because of the accumulation of toxic substances. We tested pesticides for interaction with organic-acid secretion, as indicated by inhibition of paminohippuric acid (PAH) transport. We used primary monolayer cultures of winter flounder (Pseudopleuronectes americanus) proximal tubules to measure PAH flux at 10 mu M PAH in the presence of each of several pesticides at 0.1 mM. The use of cultures allowed us to measure unidirectional fluxes in controlled electrochemical environments and to monitor tissue condition by measurement of transepithelial potential difference (PD), resistance (R), and sodium-dependent glucose transport, as indicated by phloridzin-sensitive short-circuit current (PSC). We measured PAH transport in the presence of four closely related organochlorine herbicides. (2-Methyl-4chlorophenoxy)-acetic acid (MCPA) and 2-(2-methyl-4-chlorophenoxy) propionic acid (mecoprop) at a concentration of 0.1 mM had no effect on PAH transport and were not tested further. 1.0 mM (2,4dichlorophenoxy) acetic acid (2,4-D) inhibited PAH transport by about 80%; 0.1 mM 2,4-D inhibited by 20%. 1.0 mM 2-(2,4-dichlorophenoxy) propionic acid (dichloroprop) inhibited PAH secretion by about 70% transport and 0.1 mM by 40%. Neither herbicide affected PD, R, or PSC, indicating that the effect was specific to PAH transport and not a loss of tissue integrity. Five other herbicides were tested for

6.

7.

58

Human Health inhibition of PAH transport at 0.1 mM. Of the five, only N-(phosphonomethyl) glycine (glyphosate) inhibited slightly but significantly; a concomitant decrease in PD, R, and PSC suggested that inhibition by glyphosate was the result of non-specific tissue damage. The polar metabolite of DDT 2,2-bis (pchlorophenyl) acetic acid (DDA), at concentrations of 0.001 mM and above inhibited PAH transport with a K sub (1/2) of 0.03 mM; at 0.5 mM DDA, reabsorptive flux increased to 400% of control and PSC decreased, indicating that the high DDA concentration caused tissue damage. This biomonitoring system shows that several pesticides interact with organic-anion secretion by the proximal tubule and may, through competition or direct tissue damage, increase the dwell time of toxins in the body. 8. Flaherty, D. K., C. J. Gross, K. L. McGarity, P. A. Winzenburger, and S. J. Wratten. 1991. The effect of agricultural herbicides on the function of human immunocompetent cells. II. Effect on natural killer cell and cytotoxic T cell function. In Vitro Toxicology. A Journal of Molecular and Cellular Toxicology 4 , no. 2 : 145-60. In vitro lysis of human natural killer (NK) sensitive target cells and mixed lymphocyte culture (MLC) assays were used to determine the effect of agricultural herbicides on NK cells and cytotoxic Tlymphocytes (CTL). Test chemicals ranged from 10 to 0.01 millimolar. Mononuclear cell preparations were derived from 53 human subjects and incubated with active herbicidal ingredients, active ingredients covalently bound to human serum albumin, or product formulations. The active ingredients were: propachlor (1918167), 2-chloro-N,N-diallylacetamide (93710), glyphosate (1071836), ethylparathion (56382), and methyl-parathion (298000). Conjugates of the active ingredients were prepared for use in the in vitro assays to elucidate the role of hapten protein conjugates in immunosuppression. The product formulations used were: Ramrod, Randox, and Roundup. Initial studies were conducted to determine the normalcy of the cells used in the MLC assay, the optimal concentrations of nonspecific and positive immunosuppressive agents, and the inherent toxicity of the test chemicals. Test chemical effects were compared to untreated, positive, and negative controls. Data indicated that neither parent versions of the test chemicals nor product formulations altered the in vitro growth of CTL or the function of NK cells found in blood from normal human subjects. Flaherty, D. K., and M. Panyik. 1990. The effect of agricultural chemicals and formulations on the bactericidal capacity of human peripheral blood phagocytic cells. In Vitro Toxicology. A Journal of Molecular and Cellular Toxicology 3 , no. 3 : 195-204. An investigation was conducted to determine whether changes occurred in the bactericidal capacity of human peripheral blood phagocytic cells following in vitro exposure to agricultural formulations and chemicals. The kinetic parameters of killing and the intercellular survival of the bacteria within phagocytic cells were examined after 30 and 60 minutes of incubation with Ramrod (1918167), Randox (93710), and Roundup (1071836) herbicides and parathions. There was no effect noted on the following experimental parameters: total extracellular bacterial killing, the percent of bacteria killed extracellularly per minute, the percent of bacteria ingested by the phagocytic cells, or the intracellular percent kill rate per minute. The intracellular survival of opsonized bacteria was about 2 to 4% at 30 or 60 minutes when treated or control phagocytic cells were used in the assay. The authors conclude that the chemicals have no effect on bacterial killing by human phagocytic cells following exposure to any of the test compounds. Geiger, C. P., and E. J. Calabrese. 1985. The effects of five widely used pesticides on erythrocytes of the Dorset Sheep, an animal model with low erythrocyte glucose-6-phosphate dehydrogenase (G-6-PD) activity. Journal of Environmental Science and Health Part A 20 , no. 5 : 521-27. The effects of malathion (121755), aldicarb (116063), diuron (330541), glyphosate (1071836), and metolachlor (51218952) on erythrocytes were investigated using sheep blood. Dorset-sheep erythrocytes, low in glucose-6-phosphate-dehydrogenase (G6PD) activity (comparable to human G6PD deficiency), were used. Sheep blood samples were treated with 1, 10, or 100 parts per million (ppm) of malathion, aldicarb, and metolachlor; 0.7, 7, or 70 ppm of diuron; or 0.25, 2.5, or 25 ppm of glyphosate in vitro for 2 hours at 37 degrees-C. Erythrocyte methemoglobin (METHb) and glutathione (GSH) concentrations were determined. At 100 ppm malathion and metolachlor GSH concentrations were reduced by 59 and 45 percent, respectively. Although 10ppm malathion produced a 9.9 percent decrease in GSH, this was not statistically significant. None of the pesticides displayed any effect on 59

9.

10.

Human Health METHb. The authors conclude that malathion and metolachlor can produce an oxidative stress in blood and that the other agents pose no hemolytic risk for G6PD deficient individuals. Predictive utility of Dorset-sheep erythrocytes for human G6PD deficient erythrocytes may be acceptable for some pollutants. 11. Gentile, T. J., and E. J. Calabrese. 1987. Screening for potential hemolytic responses to environmental agents using a bioactivation system: evalution of six pesticides. Journal of Environmental Science and Health, Part A: Environmental Science and Engineering A22 , no. 5 : 427-44. An in vitro microsomal enzyme system was used to evaluate the effects of aldicarb (116063) (ALD), carbaryl (63252) (CARB), diuron (330541) (DU), glyphosate (1071836) (GL), malathion (121755) (MA), and metolachlor (51218452) (MET), and the bioactivated metabolites of these pesticides on the glucose-6-phosphate-dehydrogenase (GPD) deficient erythrocytes of Dorset-sheep. The levels of reduced glutathione (GSH) in the blood and methemoglobin (MetHb) formation were used as indicators of GPD erythrocyte deficiency. None of the pesticides tested nor any of their metabolites, acted as oxidant stressors with respect to the GPD deficient erythrocytes, with the exception of DU. A significant increase in the levels of MetHb formation was recorded in the DU treated erythrocytes. Erythrocytes treated with ALD, MA and MET, and the metabolites of these products, presented significant reduction in the activity of acetylcholinesterase. The authors conclude that the lack of significant effects of MA on GSH and MetHb contradict previous findings regarding GSH depletion and MetHb formation induced by MA in sheep erythrocytes. Goldsmith, D. F. 1988. Agricultural health hazards. Western Journal of Medicine 148: 80-81. A brief overview is presented of the risks to American farm workers (farm managers; pesticide mixers, loaders and applicators; greenhouse/nursery workers) of respiratory diseases, cancer, dermatitis and accidental mortality. The top 3 pesticides cited in California pesticide illness reports are propargite, sulfur and glyphosate. Health and Safety Executive. 1987. Biological monitoring of workers exposed to organophosphorus pesticides., HMSO Publication Centre, London, United Kingdom. This is a further revision of Guidance Note MS 17, originally published in 1981 (see CIS 81798) and revised in 1986. It contains a brief roundup of information on organo-phosphorus pesticides; potential occupational sources of exposure (manufacture and packaging; transport, storage and distribution; application and use; handling used containers e.g. scrap recovery); routes of absorption; biological effects; clinical manifestations of organo-phosphorus poisoning; measurement of cholinesterase activity as an index of organo-phosphorus uptake and effect; biological monitoring; electro-physiological monitoring; emergency treatment. (United Kingdom). Hietanen, E., K. Linnainmaa, and H. Vainio. 1983. Effects of phenoxyherbicides and glyphosate on the hepatic and intestinal biotransformation activities in the rat. Acta Pharmacologica Et Toxicologica 53 , no. 2 : 103-12. The effects of 2,4-dichlorophenoxyacetic acid (94757) (2,4-D), 4-chloro-2-methylphenoxyacetic acid (94746) (MCPA), ethyl-2-(p-chlorophenoxy)-2-methyl-propionate (637070) (clofibrate), and glyphosaste (1071836) on drug metabolizing enzyme activities were examined in livers and intestines of Wistar-rats. Doses of 150 or 200 milligrams per kilogram (mg/kg) 2,4-D or MCPA were given by gavage 5 days per week for 2 weeks. Clofibrate was given at a dose of 200 mg/kg, and glyphosate was given at 500 mg/kg for 4 days, then at 300 mg/kg for the remainder of the interval. Rats were killed 24 hours after the last dose, and the liver and a segment of the proximal small intestine were removed. Liver homogenates were prepared and microsomes were isolated. Enzyme and compositional assays were performed on liver homgenates. Postmitochondrial fractions were prepared from intestinal mucosa and analyzed. Protein content of liver microsomes was unaltered after administration of the herbicides or clofibrate. No changes were seen in soluble protein content. Microsomal cholesterol content was significantly decreased in animals treated with 100 mg/kg 2,4-D. A decrease was also seen in rats treated with clofibrate or glyphosate. No changes were found in microsomal phospholipid contents. Protein content of intestinal mucosa decreased in rats given high doses of MCPA. Microsomal cytochrome-P-450 did not change significantly in the liver after

12.

13.

14.

60

Human Health administration of phenoxy acids and increased only slightly after clofibrate. Glyphosate administration resulted in a lower concentration of cytochrome-P-450 in the liver. The most marked enzyme changes were in rats treated with MCPA, in which ethoxycoumarin-O-deethylase activity increased more than 200 percent. Hepatic activity of epoxide-hydroxylase was significantly higher after treatment with 2,4-D or MCPA, which also significantly decreased glutathione-S-transferase activity. Rats treated with a high dose of MCPA showed significantly decreased uridine-diphosphate-glucuronosyl-transferase activity. The authors conclude that phenoxy acid herbicides and glyphosate have opposite effects on the activities of drug metabolizing enzymes. 15. Hui, X., R. C. Wester, P. S. Magee, and H. I. Maibach. 1995. Partitioning of chemicals from water into powdered human stratum corneum (callus): A model study. In Vitro Toxicology 8 , no. 2 : 159-67. To evaluate the bioavailability of chemicals, which cover a broad range of molecular properties and bioactive interest, the partition coefficients of 12 14C-labelled compounds between water and powdered human stratum corneum (PHSC) have been determined. Under fixed experimental conditions, the amount of chemical required to reach complete partitioning was limited and compound specific. Log PC (PHSC/w) of lipophilic compounds gave a linear correlation with log PC (o/w) (r-2 = 0.95), while those of hydrophilic compounds seemed to parallel their reverse log PC (o/w) (r-2 = 0.88). Moreover, the values of log PC (PHSC/w) for those compounds with log PC (o/w)' values between -2 and +1.5 were approximately the same. Results suggested that partitioning processes of higher lipophilic (gt +2.0) or hydrophilic chemicals (lt -2.0) are governed by the lipid or the protein domains, respectively, while those intermediate lipophilic and hydrophilic compounds are controlled by both domains, and hydration. Based on these findings, a multiple nonlinear regression was developed to predict degree of chemical partitioning into the PHSC. The PC (PHSC/w) may substitute the U.S. EPA model, PC (octanol/water) to predictive dermal exposure. It represents a stable, biologically derived system that is inexpensive, easy to use and useful for in vitro toxicology prediction. Institute of Occupational Health. 1986. Section 1: occupational health and chemical safety. Extended abstracts. International symposium on health and environment in developing countries. (Finland). International Programme on Chemical Safety (IPCS), World Health Organization, distribution and sales service. 1994. Glyphosate. Geneve, Switzerland. Data on the effects of glyphosate in humans are limited. Three irritation/sensitization studies in human volunteers indicated no effect. The herbicide formulation Roundup containing glyphosate is acutely toxic to humans when ingested intentionally or accidentally. Animal studies show that glyphosate is not carcinogenic, mutagenic or teratogenic. Detailed abstracts in French and Spanish. (Switzerland). Jamison, J. P., J. H. M. Langlands, and R. C. Lowry. 1986. Ventilatory impairment from preharvest retted flax. British Journal of Industrial Medicine 43 , no. 12 : 809-13. Flax fibres are freed from the rest of the plant by microbial degradation (retting) of dead plants. The traditional process of dew-retting involves cutting the plant and letting it lie in the field. In preharvest-retting, the weed killer glyphosate is sprayed on the crop, which then rets before cutting. The acute bronchoconstrictor responses of 11 normal subjects to dust from dew-retted and from preharvest-retted flaxes were compared in a double-blind crossover fashion. There were no significant differences in the dust levels nor in the size of the dust particles in the experimental dust room. The decreases in pulmonary function after 6 hours of dust inhalation were significantly larger after preharvest-retted flax dust than after dew-retted flax dust. The subjects also reported more symptoms after inhaling pre-harvest-retted flax dust. The acute bronchoconstrictor response to flax dust is increased by pre-harvest-retting, suggesting an increased risk of byssinosis. (United Kingdom).

16.

17.

18.

61

Human Health 19. Jauhiainen, A., K. Rasanen, R. Sarantila, J. Nuutinen, and J. Kangas. 1991. Occupational exposure of forest workers to glyphosate during brush saw spraying work. American Industrial Hygiene Association Journal 52: 61-64. The purpose of this study was to measure forest workers' exposure to the herbicide glyphosate during silvicultural clearing work done with brush saws equipped with pressurized herbicide sprayers. Both the exposed (study) group and nonexposed (control) group contained five persons who were medically examined before and after their 1-week working period (including laboratory tests) for possible health effects. In addition, exposure to glyphosate was measured in the study group from samples taken from the workers' breathing zone and from urine samples collected during the afternoons of the work week. The laboratory tests and urinary glyphosate analyses were repeated for the exposed group 3 weeks later, when the men had entirely stopped their work with the herbicide. Exposure to glyphosate through the workers' breathing zone was low. The highest value found was 15.7 .m.g/m3. In this study, a biological monitoring method was also developed to monitor the workers' exposure to glyphosate. Urine concentrations were under the gas chromatographic detection level of <0.1 ng/.m.L (<1.0 .m.mol/L). No major differences were noted, either in medical examinations or in the laboratory tests performed, between the exposed and control groups before and after the work period. (Finland). Kageuka, M., Y. Hieda, K. Hara, M. Takamoto, Y. Fukuma, and S. Kashimura. 1988. Japanese Journal of Legal Medicine 42 , no. 2 : 128-32. Analysis of glyphosate and (aminomethyl) phosphonic acid in a suspected poisoning case. (Japan). Kale, P. G., B. T. Petty, Jr., S. Walker, J. B. Ford, N. Dehkordi, S. Tarasia, B. O. Tasie, R. Kale, and Y. R. Sohni. 1995. Mutagenicity testing of nine herbicides and pesticides currently used in agriculture. Environmental and Molecular Mutagenesis 25: 148-53. Nine herbicides and pesticides were tested for their mutagenicity using the Drosophila sexlinked recessive lethal mutation assay. These are Ambush, Treflan, Blazer, Roundup, 2,4-D Amine, Crossbow, Galecron, Pramitol, and Pondmaster. All of these are in wide use at present. Unlike adult feeding and injection assays, the larvae were allowed to grow in medium with the test chemical, thereby providing long and chronic exposure to the sensitive and dividing diploid cells, i.e., mitotically active spermatogonia and sensitive spermatocytes. All chemicals induced significant numbers of mutation in at least one of the cell types tested. Some of these compounds were found to be negative in earlier studies. An explanation for the difference in results is provided. It is probable that different germ cell stages and treatment regimes are suitable for different types of chemicals. Larval treatment may still be valuable and can complement adult treatment in environmental mutagen testing. Kammerer, M. 1995. Intoxication by glyphosate based herbicides. Recueil De Medecine Veterinaire De L'Ecole D'Alfort 171: 149-52. The use of weed-killer based on glyphosate is widespread and pets are more and more likely to be exposed to it. Often presented as nondangerous, this herbicide can be toxic, after ingestion or contact. Numerous accidents in man, as well as the telephone calls received by the CNITV (National Toxicology Information Centre) allow the risk for animals to be assessed. The symptomatology concerns essentially the digestive tract and treatment of the symptoms generally leads to a favourable evolution. (France). Lavy, T. L., J. E. Cowell, J. R. Steinmetz, and J. H. Massey. 1992. Conifer seedling nursery worker exposure to glyphosate. Archives of Environmental Contamination and Toxicology 22: 6-13. This study addresses the measurements of glyphosate exposure received by 14 workers employed at two tree nurseries. The applicators, weeders, and scouts monitored all wore normal work clothing, which for applicators was a protective suit, rubber gloves and boots. Measurements were made of the glyphosate that was dislodged from conifer seedlings under water rinses taken twice weekly from May through August. Only 1 of these 78 dislodgeable residue samples were positive for glyphosate. Nine cotton gauze patches were attached to the clothing of each worker one day per week during this same period. Hand washes were taken on the same day that patches were worn. Most 62

20.

21.

22.

23.

Human Health patches and hand washes from applicators and weeders contained measurable amounts of glyphosate. Analyses of individual patches showed that the body portions receiving the highest exposure were ankles and thighs. For scouts only 1 of 23 hand washes contained glyphosate. To provide a measure of the exposure occurring via all exposure routes (dermal, ingestion, and inhalation) an analysis was made of the total urine excreted. For most workers a daily total urine collection was made for 12 consecutive weeks. Urine analysis, the biological monitoring tool used to assess the total amount absorbed via all avenues, did not reveal any positive samples. The lower limit of method validation for glyphosate in the urine samples was 0.01 .m.g/ml. High rainfall, or irrigation as needed, in conjunction with normal field dissipation avenues and worker training were cited as contributing factors for the low amounts of glyphosate exposure found. None of the exposure parameters indicated that glyphosate exposure poses a threat to human health when used under normal nursery conditions. 24. Li, A. P., and T. J. Long. 1988. An evaluation of the genotoxic potential of glyphosate. Fundamental and Applied Toxicology 10 , no. 3 : 537-46. The genotoxic potential of the herbicide glyphosate (1071836) was studied using a variety of microbial and mammalian systems. Microbial genotoxicity was assessed using the Ames Salmonella histidine reversion assay with strains (TA-1535), (TA-1537), (TA-1538), and (TA-98); the Escherichia coli (WP-2) reverse mutation test with the tryptophan-hcr strain; and a recombination assay with normal (H-17, rec(+)) and recombination deficient (M-45, rec(-)) strains of Bacillus subtilis. Mammalian genotoxicity was tested in vitro using the Chinese hamster ovary cell hypoxanthine/guaninephosphoribosyl-transferase (CHO/HGPRT) gene mutation assay and the rat primary hepatocyte DNA repair assay and in vivo using the rat bone marrow cytogenetics assay. Glyphosate at levels between 10 to 5000 micrograms per plate had no significant effect on the number of revertants in any of the Salmonella strains tested, nor did the herbicide have any significant dose response effect on either the number of (WP-2) revertants or the growth inhibition of the (M-45) and (H-17) Bacillus strains. Glyphosate at levels of 5, 17.5, and 22.5 mg/ml had no significant mutagenic effect on CHO cells at levels of S9 between 1 and 10 percent, but cytotoxic effects were associtaed with 20 and 25 mg/ml glyphosate in the absence and presence of S9, respectively. The effects of glyphosate on hepatocyte DNA repair were unremarkable, and bone marrow cytogenetics revealed no significant increases in either chromosomal aberrations or achromatic lesions in the herbicide treated animals. The authors conclude that glyphosate does not pose a genetic risk to man. Libich, S., J. C. To, R. Frank, and G. J. Sirons. 1984. Occupational exposure of herbicide applicators to herbicides used along electric power transmission line right-of-way. American Industrial Hygiene Journal 45 : 56-62. Lu, F. C., and M. L. Dourson. 1992. Safety/risk assessment of pesticides: principles, procedures and examples., Toxicology Letters 64-65 Spec. No. 783-787. The principles and procedures for the assessment of the safety/risk of chemical used by the relevant WHO and EPA expert groups were outlined. The assessment in terms of acceptable daily intakes (ADIs) and reference doses (RfDs) of 25 pesticides is listed. The pesticides assessed are acephate, alachlor, amitrole, azinphos methyl, benomyl, biphenthrin, bromophos, chlordane, chlorthalonil, cyhalothrin, DDT, EPTC, ethion, folpet, fosetyl-al, glyphosate, isofenphos, methomyl, methyl mercury, paraquat, phosphamidon, systhane, terbutyn, tribultyltin oxide, and vinclozin. In addition, their critical effects, the no-observed-effect levels and the size of the safety/uncertainty factors used are also listed to illustrate the diversity of the toxic effects and the resulting assessments. Furthermore, the enormous amount of data reviewed and the complex scientific judgement involved are also indicated. Considering the various uncertainties existing, the ADIs and RfDs do not differ appreciably in most instances. However, marked differences exist between the ADIs and RfDs of DDT and chlordane. It is suggested that re-evaluation be done on these, and perhaps other, chemicals.

25.

26.

63

Human Health 27. 28. Mack, R. B. 1993. The night the light went off in Sestos. Roundup (glyphosate) poisoning. North Carolina Medical Journal 54 , no. 1 : 35-36. Maibach, H. I. 1986. Irritation, sensitization, photoirritation and photosensitization assays with a glyphosate herbicide. Contact Dermatitis 15, no. 3: 152-56. The irritation, sensitization, photoirritation, and photosensitization potential of N(phosphonomethyl) glycine (1071836) (glyphosate) was studied in human subjects. A formulation that contained 41 percent glyphosate was applied to intact or Draize type abraided skin of 346 volunteers. Single and 21 day cumulative irritancy and modified Draize type skin irritation assays were performed. A phototoxicity and a modified photo Draize skin sensitization study were also performed. For comparison purposes, a general all purpose liquid cleaner, a baby shampoo, and a liquid dishwashing detergent were also evaluated. On unabraided skin, glyphosate showed no greater irritation potential than either the all purpose cleaner, the dishwashing liquid, or the baby shampoo. When tested on abraided skin, glyphosate had a slightly greater incidence of erythema at 24 hours; however, the 48 hour reading indicated that the irritancy potential was similar to that of the cleaner and dishwashing liquid. In the 21 day cumulative irritancy assay, glyphosate and the baby shampoo were less irritating than either the cleaner or the dishwashing liquid. No evidence of skin sensitization was seen. Glyphosate demonstrated no potential for photoirritation or photosensitization. The author concludes that his data provide a baseline for choosing appropriate diagnostic patch test concentrations. Ten percent glyphosate in water should be nonirritant. Marrs, T. C. 1993. Organophosphate poisoning. Pharmacology & Therapeutics 58 , no. 1 : 51-66. The present review discusses the structure of the anticholinesterase organophosphates (Ops), which are used predominantly as insecticides. OP poisoning can occur in a variety of situations and can be accidental or suicidial. It is common in developing countries. The cholinergic syndrome is caused by acetylcholinesterase inhibition, and diagnosis is based on the clinical signs and symptoms as well as the measurement of inhibition of erythrocyte acetylcholinesterase and/or plasma cholinesterase activity. Antidotal treatment is with atropine, an enzyme reactivator such as pralidoxime and diazepam. Anticholinesterase Ops may produce effects other than the acute cholinergic syndrome, including the intermediate syndrome. Later effects may include organophosphorus-induced delayed neuropathy. Certain Ops are exploited for their anticholinesterase effects, including defoliants such as DEF, herbicides such as glyphosate, fire retardants and industrial intermediates. The toxicology of this group is heterogeneous and they may or may not possess anticholinesterase activity. Marshall, B. E. 1990. Chemical control of water hyacinth in Lake Chivero formerly Lake McIlwaine Zimbabwe and the risk to the public. Zimbabwe Science News 24: 96-99. (Zimbabwe). Martinez, T. T., W. C. Long, and R. Hiller. 1990. Comparison of the toxicology of the herbicide Roundup by oral and pulmonary routes of exposure. Proceedings of the Western Pharmacology Society, 33:193-97. The toxicity of the herbicide Roundup (1071836) was studied. Male Sprague-Dawley rats were exposed to Roundup intratracheally or orally and examined pathologically after 24 hours. Death rates of 60% and 80% were seen following intratracheal administration of 0.2 and 0.4 milliliters (ml) of 1% Roundup, respectively. These rates increased to 80% and 100% following doses of 0.2 ml and 0.4 ml of the 18% Roundup concentrate, respectively. Significant increases in lung weights were seen following intratracheal treatment with concentrated Roundup compared with saline treated controls. Animals treated intratracheally with concentrated Roundup demonstrated severe respiratory difficulties as well as foaming edema in the lungs and atria engorged with unclotted blood. Using an oral dose of concentrated Roundup of 3 ml resulted in the death of 40% of the animals while 100% of the animals died following an oral dose of 5 ml. Delayed respiratory difficulty was seen following the administration of these doses along with edema of the stomach and gut. Few clinical or pathological changes were seen following the oral administration of 1% Roundup.

29.

30.

31.

64

Human Health 32. Menkes, D. B., W. A. Temple, and I. R. Edwards. 1991. Intentional self-poisoning with glyphosate-containing herbicides. Human Experimental Toxicology 10: 103-7. Four cases of self-poisoning with 'Roundup' herbicide are described, one of them fatal. One of the survivors had a protracted hospital stay and considerable clinical and laboratory detail is presented. Serious self-poisoning is associated with massive gastrointestinal fluid loss and renal failure. The management of such cases and the role of surfactant toxicity are discussed. (New Zealand) . Moseley, C. L., and K. Anderson. 1985. Health hazard evaluation report , HETA-83-341-1557. Bureau of Reclamation, U.S. Department of the Interior, Denver, Colorado . 22 p. Breathing zone samples were analyzed for the pesticides 2,4-D (94757), Banvel (1918009), Buctril (1689845), and Roundup (38641940) at field stations of the Bureau of Reclamation(Bureau) (SIC-9511), US Department of the Interior at Fresno, Tracy, and Red Bluff, California and Bismark, North Dakota in August, 1983 and June, 1984. The survey was requested by the Bureau to evaluate its pesticide application program. Work practices were observed during mixing, spraying, and cleanup operations. Pesticide concentrations in micrograms per cubic meter (microg/m3) were: 2,4-D, 1.8 to 2.6; Banvel, 1.2 to 1.4; Buctril, less than 2; and Roundup, less than 28. Only 2,4-D has a published standard, 10,000 microg/m3 (OSHA). Observation of work practices indicated that Bureau employees had a good understanding of the hazards associated with pesticides. The authors note that the low concentrations of active ingredients (on the order of 0.5 to 3 percent) in the pesticide formulations contributed to the low airborne concentrations measured. They conclude that in the event of an accident or spraying under windy conditions, Bureau employees would have a slight increased risk of exposure. Recommendations include employee education, using protective clothing during mixing, and implementing programs of exposure recordkeeping. Olorunsogo, O. O., E. A. Bababunmi, and O. Bassir. 1979. Effect of glyphosate on rat liver mitochondira in vivo. Bulletin of Envionmental Contamination and Toxicology 22 , no. 3 : 25764. The effects of N-phosphonomethylglycine (1071836) (glyphosate) on energy conservation in liver mitochondiria were studied in rats. Female Wistar-rats received intraperitoneally 15, 30, 60, and 120 milligrams per kilogram (mg/kg) glyphosate. Rats were killed 5 hours after treatment. Livers were excised and placed in 0.25 molar sucrose. Homogenates were centrifuged. Mitochondrial pellets were suspended in sucrose. Oxygen uptake was measured polarographically and the effect on the respiratory control ratio was calculated. Adenosine-triphosphatase (ATPase) activity was determined after reaction of mitochondrial fractions in buffer, ATP, potassium-chloride, and sucrose. Phosphate in supernatant was determined after addition of trichloroacetic-acid. The activities of isocitrate, glutamate and beta-hydroxybutyrate were determined spectrophotometrically by following the rate of formation of the reduced coenzyme nicotinamide-adenine-dinucleotide. The mictochondrial suspension was added to a reaction medium that contained sodium-succinate and potassium-ferric-cyanide, and succinatedehydrogenase activity was measured. Cytochrome-c-reductase activity and cytochrome-oxidase activity were determined. Glyphosate enhanced the rate of oxygen consumption in treated rats as compared to controls when the reaction medium was deficient in phosphate acceptance. Respiration stimulation was significantly reduced in mitochondria after glyphosate poisoning. At 30, 60, and 120 mg/kg, the reduction was 32.6, 43.8, and 46.0 percent, respectively. ATPase activities were enhanced in a dose dependent fashion in treated animals as compared to controls. The mitochondrial dehydrogenases were slightly enhanced at 60 and 230 mg/kg glyphosate. There was no significant difference in activities of the cytochrome-c-reductase and the cytochrome-oxidase systems in glyphosate treated animals as compared to controls. The authors suggest that uncoupling of mitochondrial oxidative phosphorylation may be a major lesion in glyphosate intoxication. Pilliere, F., P. Maigret, R. Garnier, P. Harry, F. Baud, and M. L. Efthymiou. 1993. Poisoning by a weedkiller preparation containing glyphosate. Presse-Med. 22 , no. 10 : 494. (France). Rank, J., A. G. Jensen, B. Skov, L. H. Pedersen, and K. Jensen. 1993. Genotoxicity testing of the herbicide Roundup and its active ingredient glyphosate isopropylamine using the mouse bone

33.

34.

35.

36.

65

Human Health marrow micronucleus test, Salmonella mutagenicity test, and Allium anaphase-telophase test. Mutation Research 300 , no. 1 : 29-36. The genotoxic potential of the herbicide Roundup and its active agent, glyphosate isopropylamine salt, was studied in three different assays. No clastogenic effects were found in the mouse bone marrow micronucleus test for either of the two agents. In the Salmonella assay only Roundup was tested. It showed a weak mutagenic effect for the concentrations 360 mu-g/plate in TA98 (without S9) and 720 mu-g/plate in TA100 (with S9). These concentrations are close to the toxic level. The anaphase telophase Allium test showed no effect for the glyphosate isopropylamine salt, but a significant increase in chromosome aberrations appeared after treatment with Roundup at concentrations of 1.44 and 2.88 mg/l when calculated as glyphosate isopropylamine. The most frequent aberrations observed could be characterized as disturbances of the spindle. 37. Sale, J. S. P., and P. M. and others Tabbush. 1986. The use of herbicides in the forest - 1986. Forestry Commission booklet 51. Edinburgh. (United Kingdom). See Biodiversity and Habitat Restoration Section. Samuel, O., L. A. Ferron, and L. St-Laurent. 1996. Evaluation of dermal exposure of a population exposed to glyphosate, and determination of a transfer coefficient for dislodgeable foliar residues.). Ministre des Ressources naturaelles du Qubec, Direction de lenvironnement forestier, Sainte-Foy, Centre de toxicologie du Qubec. 56 p. A field study was undertaken in order to better characterize glyphosate exposure risks for a population either wildberry picking or hiking in the forest. The study also aimed at establishing the relationship between measured residue levels and the amounts of either dislodgeable foliar residues or dislodgeable wildberry (raspberry) residues. Results allowed a re-evaluation of the daily doses absorbed dermally, in the various scenarios considered within the Quebec Ministry of Natural Resources impact study on the various methods of forest tending. Dermal pads were used to evaluate dermal exposure related to the various activities; these allowed better estimation of the anatomical distribution of dermal exposure (forearm, hand, head, neck). Leaf samples were collected using a leaf-punch. Average foliar and fruit dislodgeable residues were determined, and the influence of rain events of these residue levels was documented. Results indicate that dislodgeable foliar residues are generally stable in the absence of rain events in the week following application. Also, the highest dislodgeable foliar residue levels measured in this study were lower by more than a factor of two, than the levels estimated in the MNRs impact study. In the case of dislodgeable residues from raspberries, a tendency towards an increase in residue levels was noted at day 5; this seems to confirm MNRs hypothesis that part of the glyphosate absorbed by raspberry plants is migrating towards the fruits in the days following application. Results from a field study where glyphosate had been added only to raspberry fruits indicated a rapid decrease in dislodgeable residues in the period between 30 minutes and 24 hours after application. Results of dermal exposure indicate that the lower portions of the legs, the thighs and hands are most exposed during berry picking or hiking. Also, the highest dermal exposure level measured for the whole body was much lower than the levels estimated in the impact study. Using these field results, the levels calculated for other types of activities (residents, hunters and anglers) are also much lower than the ones estimated in the impact study. As a good correlation was observed in the study between dislodgeable foliar residue levels and dermal exposure levels, an average transfer coefficient was calculated for berry picking and hiking. This indicated that the 5 000 cm 2/h coefficient used in the impact study was both safe and appropriate. Samuel, O., and J. G. Guillot. 1989. tude sur lefficacit du lavage leau froide pour liminer le glyphosate sur les vtements de travail. Ministre de Lnergie et des Ressources, Centre Hospitalier de L'Universit Laval, Centre de Toxicologie du Qubec. ER89-1196. 24 p. The present study was undertaken in order to evaluate the effectiveness of washing with cold water to eliminate glyphosate residues adsorbed on working garments and thus to estimate the possibility of using fabric clothing in the place of rainwear which is less comfortable for the worker.

38.

39.

66

Human Health Four experiments were performed on working trousers using both technical glyphosate and the commercial formulation VISION. The total quantity of glyphosate deposited on each garment was 22 !4 mg. Two other experiments were also performed on overalls contaminated with 545 and 1090 mg respectively of commercial formulation. During these experiments, the glyphosate concentrations were determined in the wash and rinse water and also on the contaminated pieces of fabric. For the last two experiments, glyphosate concentrations were measured in the soaking water. All washes were carried out with cold water in a washing machine. The results indicate that glyphosate was quickly removed from the clothing, and generally, was not detected in the rinse water. Subsequent measurement of residual glyphosate on garments show that the effectiveness of washing in cold water is greater than 99%. As appears from this study, cold water washing or a prolonged soaking efficiently removes glyphosate from the work garment. 40. Samuel, O., L. Houde, and D. Phaneuf. 1994. valuation des risques la sant humaine attribuables lutilisation de glyphosate en milieu forestier, Ministre des Ressources naturelles du Qubec, Direction de lenvironment foresteir , Sainte-Foy, Centre de Toxicologie du Qubec. RN95-3029. 72 p. At the request of the ministre des Ressources naturelles (MRN), the Centre de Toxicologie du Qubec (CTQ) assessed the potential risks to human health associated with the use of the phytocide VisionTM during forest regeneration maintenance operations. The target groups analysed were forestry workers, individuals living close to the sites treated, and forest users, including hunters and fishermen. The analysis was concerned mainly with glyphosate, which is the active ingredient in VisionTM. Where data were available, we also performed a qualitative or semi-quantitative assessment of the risks associated with the presence of additives or impurities in the commercial formulation. The health risks for workers using ground and aerial application methods were assessed on the basis of the results of a CTQ study, together with exposure data taken from the literature. Worker health risks in accidental exposure scenarios were also estimated. In the case of manual application methods, we were able to make only a qualitative assessment of the risks because of the lack of exposure data. It was based on the technical features of the application methods used. The potential public health risks were assessed by comparing the total estimated exposure dose for the various groups with the reference dose (RfD) established from experimental studies. The chronic and acute effects on the population were assessed on the basis of long-term, short-term and accidental exposure scenarios. The worst realistic case approach was used to estimate the highest exposure likely to be experienced by the population. Exposure doses were estimated on the basis of residual glyphosate concentrations measured or estimated by the MRN. Long-term exposure was assessed on the basis of scenarios which included simulations of the activities of a resident living close to a treatment site and a hunter-fisherman partly dependent on produce from the treatment site for food. The data currently available show that, in general, worker exposure is low if the work is performed in compliance with the recommended safety measures, and it seems unlikely that worker health suffers as a result. However, in the accidental exposure scenarios, dermal irritation may occur if the exposed worker is unable to wash immediately. In the population exposure scenarios, the estimated long-term exposure doses for residents and hunter-fishermen were below the reference dose, despite our conservative assumptions. The results obtained suggest that the use of glyphosate involves only negligible risks in terms of systemic effects and effects on reproduction and development. The risk of acute toxicity is also negligible in the case of short-term exposure. In particular, there is only a very slight risk involved in consuming sprayed raspberries or water from the treatment site on the day of treatment. The amount that would have to be consumed to reach the subchronic reference does RfD is enormous, and is unlikely to occur in reality. According to the assumptions put forward, dermal exposure has the greatest impact in terms of total estimated dose. However, since a number of uncertainties are present in the dermal exposure estimates more knowledge of dermal exposure values is needed before the potential danger can be estimated more accurately. The accidental population exposure scenario suggests that a person sprayed accidentally is unlikely to suffer any systemic toxicity problems. However, since Visions commercial formulation is a

67

Human Health potential irritant, dermal irritation may occur in the case of accidental spraying or extended contact with newly-treated vegetation. Our estimates suggest that workers who comply with the recommended safety measures are unlikely to suffer systemic toxicity problems. Moreover, glyphosate used in forest operations should not increase the public exposure dose to a health-threatening level. However, as a number of uncertainties were raised during the exercise, the MRN should continue to inform the population of the place and date of phytocide applications, and state that wild fruits should not be consumed immediately afterwards. It would also be useful to specify dermal exposure values, especially absorption rates, and to harmonize environmental sampling strategies so as to reduce the uncertainties inherent in the process of estimating exposure. 41. Sawada, Y., Y. Nagei, M. Ueyama, and I. Yamamoto. 1988. Probable toxicity of surface-active agent in commercial herbicide containing glyphosate. Lancet I no. 8580 : 299. Between June 1984 and March 1986 there were 56 cases of toxic reactions to the chemical Roundup (1071836), a commercial herbicide containing glyphosate (1071836). Clinical findings in these cases suggested that it was the surface active agent in Roundup, polyoxyethyleneamine, rather than glyphosate, which caused the toxic reactions. The mean amount of Roundup ingested in these cases was about 120 milliliters. Of the 56 cases, 48 were attempted suicides and three were accidents. Gastrointestinal symptoms included sore throat, abdominal pain, and vomiting. Erosion of the pharynx, esophagus, and stomach were evident on endoscopy. Autopsy revealed erosion, necrosis, and hemorrhage of the jejunum and ileum. Cardiovascular symptoms included oliguria, anuria, and hypotension in all fatal cases and transiently in survivors. Central nervous system symptoms were present in some cases. The clinical picture was one of hypovolemic shock. Polyoxyethyleneamine has been found to have a median lethal dose less than 1/3 that of glyphosaste, and this class of surfactants has been found to cause gastrointestinal and central nervous system symptoms and hemolysis. Two patients who ingested a surface active agent showed strikingly similar clinical pictures. Shuttleworth, W. A., and J. N. Evans. 1994. Site-directed mutagenesis and NMR studies of histidine-385 mutants of 5-enolpyruvylshikimate-3-phosphate synthase. Biochemistry 33 , no. 23 : 7062-68. The site-directed mutagenesis of His-385 of 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase is reported. The steady-state kinetics for two mutants, H385Q and H385A, are compared with that of the wild-type enzyme. H385Q EPSP synthase was found to have 25% wild-type enzyme activity, whereas H385A EPSP synthase retained 1% activity. The KM values for Pi and shikimate 3phosphate were unaffected, whereas the KM for phosphoenolpyruvate (PEP) was increased 10 times for H385Q EPSP synthase. The KM for EPSP was unaffected in H385Q but raised by a factor of 10 in H385A EPSP synthase. The binding of glyphosate was studied by fluorescence spectroscopy and by 31 P NMR spectroscopy. Direct observation of the enzyme-intermediate complexes by 13C NMR spectroscopy with [2,3-13C]phosphoenolpyruvate was studied for the mutant enzymes and compared with the wild type. Under equilibrium conditions, H385A EPSP synthase does not accumulate enzymebound EPSP. These results suggest that, while critically located in the PEP binding site, His-385 is not the residue responsible for initiating catalysis through the protonation of PEP. Spickett, J. T., P. J. Dolin, M. R. Phillips, and C. J. Priestley. 1989. Patterns of pesticide usage by cereal crop farmers in Western Australia. Asia-Pacific Journal of Public Health 3 , no. 3 : 242-48. In Western Australia there has been an increase in the use of herbicides in recent years due to a change in farming practices. This change, together with more general public concern over exposure to chemicals, has resulted in farmers expressing concern over the possible long term health effects from exposure to herbicides. As part of a long term study of the possible health effects from such exposure, a survey was carried out to establish the extent of pesticide use within the cereal farming community of Western Australia. Of the 9,408 properties surveyed, 2,921 responses were received which represents a 32.2% response rate. The results indicate that a wide range of chemicals are used as insecticides, fumigants, seed dressings, seed pickles, herbicides, and rodent poisons. At the time of the survey in 1985, products containing prespruf and 1,1,1-trichloro-2,2-bis (p-chlorophenyl) ethane

42.

43.

68

Human Health (DDT) were the most popular insecticide, and products containing diquat, diclofop-methyl, chlorsulfuron and glyphosate as active ingredients represented the four most popular herbicides. (Australia). 44. Stockard, J. D. 1996. Restoration of Wingham brush 1980-1996. Eleventh Australian Weeds Conference Proceedings . See Biodiversity and Habitat Restoration Section. Talbot, A. R., Z. L. Chen, T. S. Goo, J. S. Huang, S. H. Wang, M. H. Shiaw, and S. F. Yang. 1990. Plasma levels and hemodynamics in acute glyphosate poisoning. Veterinary and Human Toxicology 32: 370. Talbot, A. R., M-H. Shiaw, J-S. Huang, S-F. Yang, T-S. Goo, S-H. Wang, C-L. Chen, and T. R. Sanford. 1991. Acute poisoning with a glyphosate surfactant herbicide Round-up. A review of 93 cases. Human Experimental Toxicology 10: 1-8. Between 1 January 1980, and 30 September 1989, 93 cases of exposure to herbicides containing glyphosate and surfactant ('Roundup') were treated at Changhua Christian Hospital. The average amount of the 41% solution of glyphosate herbicide ingested by non-survivors was 184 70 ml (range 85-200 ml), but much larger amounts (500 ml) were reported to have been ingested by some patients and only resulted in mild to moderate symptomatology. Accidental exposure was asymptomatic after dermal contact with spray (six cases), while mild oral discomfort occurred after accidental ingestion (13 cases). Intentional ingestion (80 cases) resulted in erosion of the gastrointestinal tract (66%), seen as sore throat (43%), dysphagia (31%), and gastrointestinal haemorrhage (8%). Other organs were affected less often (non-specific leucocytosis 65%, lung 23%, liver 19%, cardiovascular 18%, kidney 14%, and CNS 12%). There were seven deaths, all of which occurred within hours of ingestion, two before the patient arrived at the hospital. Deaths following ingestion of 'Roundup' alone were due to a syndrome that involved hypotension, unresponsive to intravenous fluids or vasopressor drugs, and sometimes pulmonary edema, in the presence of normal central venous pressure. (Taiwan). Temple, W. A., and N. A. Smith. 1992. Glyphosate herbicide poisoning experience in New Zealand. New Zealand Medical Journal 105 , no. 933 : 173-74. (New Zealand). Tominack, R. L., G. Y. Yang, W. J. Tsai, H. M. Chung, and J. F. Deng. 1991. Taiwan National Poison Center survey of glyphosate-surfactant herbicide ingestions. Journal of Toxicology Clinical Toxicology 29 , no. 1 : 91-109. Between January, 1986 and September 1988, the Taiwan National Poison Center recorded 97 telephone consultations (49 male, 48 female) on cases of ingestion of glyphosate-surfactant herbicide concentrate containing the isopropylamine salt of glyphosate (N-phosphonomethyl glycine, CAS 107183-6) and a non-ionic tallow amine surfactant. Eleven of the cases resulted in fatalities, all among those attempting suicide. The average amount ingested by survivors was 120 +/- 112 mL and by nonsurvivors was 263 +/- 100 mL (p less than or equal to 0.0001). The average age of survivors was 35 +/- 15 years compared to 54 +/- 11 years for fatalities (p less than or equal to 0.0002). Irritation of the oral mucous membrane and gastrointestinal tract was the most frequently reported effect. Other effects recorded were pulmonary dysfunction, oliguria, metabolic acidosis, hypotension, leukocytosis and fever. Fourteen patients received either atropine or pralidoxime plus atropine despite the fact that glyphosate does not inhibit acetylcholinesterase. Thirteen percent of patients received a urine test for paraquat or treatment customarily used for paraquat ingestion, possibly reflecting similar initial presentations following ingestion of these two herbicides. Laboratory differentiation is essential if any doubt exists about which herbicide was ingested. Patients ingesting large volumes of concentrated glyphosate-surfactant herbicide formulations require close observation and supportive treatment. (Taiwan).

45.

46.

47.

48.

69

Human Health 49. Vigfusson, N. V., and E. R. Vyse. 1980. The effect of the pesticides, Dexon, Captan and Roundup, on sister-chromatid exchanges in human lymphocytes in vitro. Mutation Research 79 , no. 11 : 53-57. The effects of the pesticides Dexon (140567), Captan (133062), and Roundup (1071836) on the sister chromatid exchanges (SCE) in human lymphocytes in vitro were evaluated. Two subjects were chosen as regular donors of blood cells. To each 5 milliliter (ml) human lymphocyte culture was added either Dexon in concentrations ranging from 2.5 to 250 milligrams per milliliter (mg/ml), Captan at 3.0 to 300 mg/ml, or Roundup at 0.25 to 25.0 mg/ml. Ethyl-methane-sulfonate (62500) (EMS) was used as positive control. For each sample and concentration, 50 well spread and differentially stained metaphases were analyzed for SCE frequency from each subject. Dexon produced a significantly higher amount of SCE than EMS at each of the concentrations tested although the response was dose related. Captan at the lowest concentration significantly increased SCEs compared to controls and EMS at the same concentration. The herbicide Roundup had the least effect on SCE, requiring the use of much higher concentrations to produce an effect. Limited data were obtained with Captan because of the toxic levels of either the fungicide or solvent used. Based on the SCE test, the authors suggest that the herbicide Roundup is at most weakly mutagenic, that Dexon is a potent mutagen and inducer of SCE, and that Captan is a significant inducer of SCE compared to the mutagen EMS. Weinstein, S. 1984. Fruits of your labour: A guide to pesticide hazards for California field workers. University of California, Berkeley. Wester, R. C., J. Melendres, R. Sarason, J. McMaster, and H. L. Maibach. 1991. Glyphosate skin binding, absorption, residual tissue distribution, and skin decontamination. Fundamental and Applied Toxicology 16 , no. 4 : 725-43. The percutaneous absorption of glyphosate (1071836) was studied in vivo and in vitro. The ability of Roundup, a commercial glyphosate formulation, applied neat and in 1:20 to 1:32 dilutions to penetrate human thigh skin samples obtained at autopsy was evaluated using flow through cells containing human plasma as the receptor fluid. The ability of carbon 14 (C-14) labeled Roundup and the 1:20 and 1:32 dilutions to bind to powdered human stratum corneum was investigated. Adult female rhesus-monkeys were administred 500 or 5400 micrograms (microg) per 200 square centimenters (cm2) labeled glyphosate topically or 9 or 93 microg glyphosate intravenously. Blood and urine samples were collected starting 24 hours before dosing and up to 8 days post dosing and assayed for C-14 activity. Selected monkeys were killed 7 days after topical exposure to determine the tissue distribution of glyphosate derived C-14 activity. Other monkeys were topically administered a 1:20 dilution of C-14 labeled glyphosate. The application sites were washed with soap and water or water 0 to 24 hours later to assess the ability of these treatments to remove glyphosate. In vitro, less than 2% of the applied glyphosate penetrated human skin. Glyphosate as Roundup or in diluted form did not bind to powdered stratum corneum. Around 95 to 99% of intravenously administered glyphosate was excreted in the urine, mostly within the first 24 hours. Following topical application only 2.2% of the 5400 microg/200 cm2 dose and 0.8% of the 500 microg/200 cm2 dose were excreted in the urine over 8 days. Based on the intravenous data, 0.8 to 2.2% of the applied doses was estimated to have been absorbed. Glyphosate was detected in the blood after intravenous administration, but not topical application. No glyphosate derived radioactivity was detected in any internal organs after topical application. Soap and water, or water removed 89.6 and 83.6% of the applied dose, respectively, 12 hours after treatment. Both treatments removed about 50% of the applied dose 24 hours after exposure. The authors conclude that the amounts of glyphosate absorbed through the skin of rhesusmonkeys is low, on the order of 0.8 to 2.2%. Since the rhesus-monkey is a good model for percutaneous absorption relevant to humans, glyphosate should have little dermal toxicity for humans. Yamano, T., and S. Morita. 1993. Effects of pesticides on isolated rat hepatocytes, mitochondria and microsomes. Archives of Environmental Contamination and Toxicology 25 , no. 2 : 271-78. The effects of selected pesticides on hepatocyte glutathione content, viability, and lipid peroxidation and hepatic mitochondrial respiration and microsome function were studied in vitro six. Twenty seven pesticides including six carbamates, seven organophosphates, six pyrethroids, halogenated compounds, and heterocycles were selected because their residues had been detected in foods at significant concentrations. Pesticides were incubated with hepatocytes, mitochondria, and

50. 51.

52.

70

Human Health microsomes isolated from the livers of male Sprague-Dawley-rats for 45 or 90 minutes. The effects on hepatocellular glutathione and ATP concentration, and lipid peroxidation, mitochondrial respiration, and microsomal NADPH/ferous ion (Fe+2) dependent lipid peroxidation were determined. Reactivity of 1 millimolar (mM) pesticide with 1mM glutathione in a phosphate buffer was investigated. Nineteen pesticides at 10 (-3)M reduced hepatocellular glutathione concentrations to below 80% of the control value after 90 minutes incubation. Oxamyl (23135220), edifenphos (17109498), methyl-parathion (298000), trichloroton (52686), permethrin (52645531), ethoxyquin (91532), pyrethrin-I (121211), and pyrethrin-II (121299) at 10(-4) significantly reduced hepatocellular glutathione concentrations. Only dichlofluanid (1085989) simultaneously reduced hepatocellular glutathione content and reacted with buffered glutathione. Oxamyl, edifenphos, methyl-parathion, and dichlofluanid induced hepatocellular lipid peroxidation. Chlorobenzilate (510156), edifenphos, dichlofluanid, and chinomethionat (2439012) caused cell death which was preceded by intracellular ATP depletion. Chloropropham (101212), cypermethrin (52315078), and clofentezine (74115245) uncoupled mitochondrial respiration. Chlorobenzilate, amitraz (33089611), chlorpyrifos (2921882), edifenphos, dichlofluanid, pyrethrin-I, and pyrethrin-II inhibited mitochondrial respiration. Diflubenzuron (35367385), chlorpyrifos, edifenphos, fensulfothion (115902), glyphosate (1071836), vamidothion (2275232), and daminozide (1596845) enhanced NADPH/Fe+2 dependent microsomal lipid peroxidation. Chloropropham, amitraz, pyrethrinI, pyrethrin-II, chinomethionat, clofentezine, and ethoxyquin inhibited NADPH/Fe+2 dependent lipid peroxidation. The authors conclude that many pesticides which leave residues in food exert potentially adverse effects on hepatic cells and organelles. 53. Yousef, M. I., K. Bertheussen, H. Z. Ibrahim, S. Helmi, M. A. Seehy, and M. H. Salem. 1996. A sensitive sperm-motility test for the assessment of cytotoxic effect of pesticides. Journal of Environmental Science and Health Part B: Pesticides, Food Contaminants, and Agricultural Wastes 31 , no. 1 : 99-115 . A sensitive sperm-motility test for the evaluation of cytotoxic effects of carbofuran and glyphosate in a defined protein-free culture medium is described. The sperm motility was compared to that obtained with a protein-containing medium. The use of protein-free medium considerably increased the sensitivity of sperm cells from rabbit and human to the toxic effects of the pesticide. The respective IC-50 values (the concentation needed to cause 50% inhibition of sperm motility) in proteinfree medium of carbofuran and glyphosate were 321 and 48.2 mu-M with human sperm, and 116 and 23.5 mu-M with rabbit sperm. Whereas, the corresponding values in protein-containing medium were 920 and 740 mu-M, and 910 and 500 mu-M with human and rabbit sperm, respectively. Our results show that testing human and rabbit sperm in protein-free medium proves to be a more sensitive method than that in protein-containing medium. Additionally, the use of rabbit sperm is a more sensitive test system than human sperm. This study suggests that the rabbit sperm test appears to have a potential for the assessment of toxicity on human reproduction. (Egypt) Yousef, M. I., M. H. Salem, H. Z. Ibrahim, S. Helmi, M. A. Seehy, and K. Bertheussen. 1995. Toxic effects of carbofuran and glyphosate on semen characteristics in rabbits. Journal of Environmental Science and Health-B 30 , no. 4 : 513-34. The present study was undertaken to investigate the effect of chronic treatment with two sublethal doses of Carbofuran (carbamate insecticide) and glyphosate (organophosphorus herbicide) on body weight and semen characteritics in male New Zealand white rabbits. Pesticide treatment resulted in a decline in body weight, libido, ejaculate volume, sperm concentration, semen initial fructose and semen osmolality. This was accompanied with increases in the abnormal and dead sperm and semen methylene blue reduction time. The hazardous effect of these pesticides on semen quality continued during the recovery period, and was dose-dependent. These effects on sperm quality may be due to the direct cytotoxic effects of these pesticides on spermatogenesis and/or indirectly via hypothalami-pituitary-testis axis which control the reproductive efficiency. (Egypt).

54.

71

Mammals

Mammals
1. Anthony, A. G., and M. L. Morrison. 1985. Influence of glyphosate herbicide on small mammal populations in western Oregon. Northwest Science 59: 159-68. The effect of glyphosate application on vegetation and small mammal populations in the Coast Range of western Oregon was investigated. Diversity, abundance, and biomass of small mammal populations increased one year post spray on glyphosate-treated sites as compared to control sites. These changes were ephemeral and above parameters were similar to prespray values two years after glyphosate application. The changes in diversity, abundance, and biomass were primarily a result of the increase in numbers of Microtus oregoni following an increase in grass and forb cover on treated grids one year postspray. The temporary effects of glyphosate treatment on vegetation had no detrimental effects on small mammal populations. Balfour, P. 1989. Effects of Forest Herbicides on Some Important Wildlife Forage Species. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, Victoria, B.C. FRDA Report 020. 58 p. (*) Seventy-one percent of the herbicide used in British Columbia forests in 1985 was glyphosate. This percentage is projected to be 78% in 1986. Therefore, glyphosate impacts on wildlife are of primary concern. Approximately 400 out of 1300 entries analyzed were glyphosate trials. Glyphosate is a broad spectrum, relatively non-selective herbicide. Indeed most of the species studied, with the exception of evergreen species, showed instances of severe damage. However, the degree and timing of severe results were varied. Being foliar active, the degree of damage is related to the amount of chemical absorbed and subsequently translocated in the plant. This is a function of method of application, plant characteristics and environmental conditions. The cut-surface applications were severely damaging in nearly all cases. During the data analysis, conflicting results with regard to season of application were found. It is often assumed that early foliar glyphosate applications will be most damaging as this is the period of most active translocation. Alternatively, when the leaves are at maximum growth there is more surface and increased absorption capabilities. The data analyzed in this study range over every possible permutation of these two theories. However, there is suggestion of some forage species showing particular trends, for instance mallow ninebark was undamaged by early treatments but severely damaged by fall treatments. Further research directed at particular seasonal/phenological sensitivities is warranted for glyphosate. Plant species may also respond on an individual basis to their general health or other factors unrelated to their species physiology. Plant vigour may also determine ability to absorb and translocate chemicals. Bancroft, B., and P. Comeau. 1990. Glyphosate and willow in moose winter range in the SBS zone. Forestry Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. Research Memo No. 150. 1 p. Significant reductions in growth and vigor of willow follows application of 2 to 3 kg a.i./ha of glyphosate from early July to late August. Whether these reductions are critical for ungulate browse has not been determined. Glyphosate at 1 kg a.i./ha does not significantly reduce willow vigor. If glyphosate is applied after leaves start to yellow (when the abscission layer is forming), it will not significantly affect willow. There appears to be no significant difference between assessed willow species susceptibility to glyphosate. However, within a species, the response is quite variable. Braathe, P. 1978. Damage by elk browsing must be taken seriously. Will our forests contain less pine in future? Norsk Skogbruk 241: 12-13. Although the introduction of evenaged forestry, with consequent large regeneration areas, is probably not the only cause of the population explosion of elks (Alces alces) in recent years, such areas do provide rich browse for elks which are also the only browsers of young pine (Pinus sylvestris). In trials in which insufficient regeneration of Picea abies was supplemented by pine, elk damage was so severe that this method had to be abandoned. Pine is also heavily attacked in natural mixed stands (pine and spruce) and in pure stands. In mixed broadleaf and conifer regeneration elk browse 72

2.

3.

4.

Mammals particularly heavily on aspen (Populus tremula) and maple (Acer spp.). It is suggested that by judicious cleaning of regeneration areas with glyphosate the food supply for elk might be greatly reduced thereby reducing the present intolerably high population density. (Norway). 5. Brewster, D. W., J. J. Warren, and W. E. Hopkins II. 1991. Metabolism of glyphosate in SpragueDawley rats: Tissue distribution, identification, and quantitation of glyphosate-derived materials following a single oral dose. Fundamental and Applied Toxicology 17: 43-51. Five groups of male Sprague-Dawley rats were orally administered a mixture of [14C]- and 12 [ C]-glyphosate (N-phosphonomethylglycine) at a dose level of 10 mg/kg body weight. The majority of radioactivity 2 hr after administration was associated with the gastrointestinal contents and small intestinal tissue. Approximately 35-40% of the administered dose was absorbed from the gastrointestinal tract, and urine and feces were equally important routes of elimination. The total body burden 7 days after administration was approximately 1% of the administered dose and was primarily associated with the bone. Total recovery for this study ranged from 95 to 102% of the administered dose. Metabolic profiles of tissues containing greater than 1% of the administered dose at various times after administration indicated that nearly 100% of the body burden of radioactivity was present as unmetabolized parent glyphosate. A minor component constituting <0.1% of the administered dose (<0.4 ppm) was observed in colon tissue from animals 2 hr after the administration of glyphosate and was also present in the GI contents of one animal 28 hr after administration of the radiolabel. The retention time for this metabolite was similar, but not identical, to the retention time for AMPA (aminomethylphosphonic acid), the major bacterial metabolite of glyphosate found in soil. Tissue extraction efficiency was always greater than 90% and stability assays indicated no significant effect of storage on either parent glyphosate or AMPA. The results from this study indicate that virtually no toxic metabolites of glyphosate were produced since there was little evidence of metabolism and essentially 100% of the body burden was parent compound with no significant persistence of material. Campbell, D. L., J. Evans, G. D. Lindsey, and W. E. Dusenberry. 1981. Acceptance by blacktailed deer of foliage treated with herbicides. USDA Forest Service, Pacific Northwest Forest and Range Experiment Station; Portland, OR: USDI, Fish and Wildlife Service, Forest-Animal Damage Control Research Project. Olympia, WA. Res. Pap. PNW-290. 31 p. To test their acceptance of foliage treated with herbicides, captive black-tailed deer were exposed to Douglas-fir seedlings and salal treated with standard formulations of 2,4,5-T, 2,4-D, atrazine, dalapon, fosamine, and glyphosate herbicides. Carriers were diesel oil and water. Tests were made from November 1977 through February 1978. Deer readily browsed 2,4,5-T treatments and most formulations of 2,4-D in oil compared with oil alone, but showed rejection of some phytotoxic glyphosate treatments. Consumption of herbicide-treated foliage did not cause noticeable health problems in test animals. Cole, E. C., W. C. McComb, M. Newton, C. L. Chambers, and J. P. Leeming. 1995. Response of small mammal and amphibian capture rates to clearcutting, burning, and glyphosate application in the Oregon coast range. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R. E. Gaskin, and J. A. Zabkiewicz (compilers). FRI Bulletin No. 192, pp. 155-57. Small mammal and amphibian capture rates were monitored on three replicate sites scheduled for hardwood conversion for one year prior to clearcutting and two years post-clearcutting. Of the 6 small mammal species analyzed, deer mice (Peromyscus maniculatus) were unaffected, creeping vole (Microtus oregoni) and vagrant shrew (Sorex vagrans) initial capture rates increased in upland areas, Townsend's chipmunk (Tamias townsendii) capture rates increased in buffered areas of logged units, and Pacific shrews (S. pacificus) and Trowbridge's shrews (S. trowbridgii) decreased in upland areas after clearcutting and burning. Capture rates of ensatina salamanders (Ensatina eschscholtzi) and Pacific giant salamanders (Dicamptodon ensatus) decreased after logging. Treatment effects on Dunn's salamanders (Plethodon dunni), rough-skin newts (Taricha grandulosa), and red-legged frogs (Rana aurora) were not detectable. Western redbacked salamander (P. vehiculum) capture rates increased in upland areas the first year after logging, but decreased to pretreatment levels the second year after logging. For all species, changes in capture rates on areas sprayed with glyphosate did not differ from rates on clearcut, unsprayed areas. 73

6.

7.

Mammals 8. Cole, E.C., W. C. McComb, M. Newton, J. P. Leeming, and C. L. Chambers. 1997. Response of small mammals to clearcutting, burning, and glyphosate application in the Oregon coast range. Journal of Wildlife Management (submitted) . We sampled small mammals on 3 replicate red alder (Alnus rubra) sites 1 year before and 2 years after the following treatments were each applied to 1 unit on each site: (1) control (uncut), (2) clearcut and broadcast burned, and (3) clearcut and broadcast burned and then sprayed with glyphosate herbicide. All sites included uncut riparian buffer strips at least 20 m wide. In upslope areas, capture rates of creeping voles (Microtus oregoni) and vagrant shrews (Sorex vagrans) increased after logging, while capture rates of Pacific shrews (S. pacificus) and Trowbridge's shrews (S. trowbridgii) decreased. Townsend's chipmunk (Tamias townsendii) capture rates increased in the buffer areas after logging. Capture rates for deer mice (Peromyscus maniculatus) did not differ among treatments. Capture rates of the shrew species that declined did not decrease in the buffer areas, indicating that riparian buffer strips may support populations of these species for up to 2 years after logging. Changes in capture rates for the 6 species analyzed on areas sprayed with glyphosate did not differ from changes on logged, unsprayed areas. Connor, J. F. 1992. "Impacts of the herbicide glyphosate on moose browse and moose use of four paired treated-control cutovers near Thunder Bay, Ontario. " M.Sc. Thesis, School of Forestry, Lakehead University. Thunder Bay, Ontario. Re-assessment of the aerial and ground observations on four paired, glyphosate treated and control, cutovers near Thunder Bay, Ontario, indicated that aerial tending with glyphosate altered the use of these cutovers by moose. The number of pellet groups favoured the control areas (p < 0.05) by 1.5 times. Additionally, the number of moose tracks and moose track aggregates were more prevalent (p < 0.05) on the controls for 2 to 3 years after treatment. Prespray data on 2 areas suggested use shifted away from glyphosate treated areas. Browse availability was significantly greater (p < 0.05) on the control plots by 18 times in the highest height class measured (201 - 350 cm), 5 times in the next highest (101 - 200 cm) but not statistically significant (p > 0.05) in the lowest (51 - 100 cm), 2 years after treatment. Due to too few replications, differences in availability 1 year after treatment were not statistically significant. Biomass of browse removed by moose was 3 to 7 times greater on controls but again these differences were not statistically significant. The average length of moose trails observed in the snow was shorter (p < 0.05) on the controls suggesting less travel time. The size (area) of moose track aggregates was the same (p > 0.05) between treatments indicating equal search time while browsing. A carrying capacity model indicated that if all cutovers were sprayed, the treatment would have a negative impact on moose densities. Glyphosate treatments should be dispersed to create a mosaic of glyphosate treated areas next to non-treated areas. Similarly, areas of seasonal importance such as aquatics, salt licks, and calving areas should have at least a non-sprayed buffer beside them if the adjacent cut area must be treated with glyphosate. Connor, J. F., and L. M. McMillan. 1990. Winter utilization by moose of glyphosate-treated cutovers. Ontario Ministry of Natural Resources, Northwestern Ontario Forest Technology Development Unit. Technical Report No. 56. 27 p. Glyphosate [N-(phosphonomethyl)glycine] is an important silvicultural tool used in the boreal forest. This study was undertaken to determine if the use of this herbicide for controlling competing shrubs in plantations is significantly reducing forage resources and subsequent overwinter utilization by moose (Alces alces) up to three years post-spray. Observations were carried out on four glyphosatetreated and control paired cutovers near Thunder Bay, Ontario. The numbers of overwinter moose tracks were not significantly different (P < 0.05) at 7 and 19 months post-spray, but they indicated a preference for the non-sprayed control areas (P < 0.05) at 31 and 43 months post-spray. The number of moose track aggregates were similar in all control and treated cutovers (P < 0.05), 7 months postspray, but were more numerous (P < 0.05) on control portions, 19, 31, and 43 months after treatment. Total track aggregate area and average track aggregate size were similar (P < 0.05) at 7, 19, and 31 months post-treatment, but total track aggregate area was significantly greater (P < 0.05) on the controls at 43 months. Available moose browse, on control areas was four times greater, and browse utilized was 32 times greater, than that in treated areas (P < 0.05) at 21 months post-spray. At 33 months post-spray, available moose browse on untreated areas was just over three times greater than on the treated areas, while utilized browse was about four and one-half times greater on controls than 74

9.

10.

Mammals on treated areas. Estimated winter moose presence, calculated from pellet counts, was two times greater (0.50 vs 0.28 moose/km2) on untreated than treated areas after 21 months (P < 0.05) and similar at 33 months post-spray. 11. . 1990. Winter utilization by moose of glyphosate-treated cutovers. Alces 26: 91-103. Glyphosate (N-(phosphonomethyl)glycine) is an important silvicultural tool used in the boreal forest. This study was undertaken to determine if the use of this herbicide for controlling competing shrubs in plantations is significantly reducing forage resources and subsequent overwinter utilization by moose (Alces alces) up to 3 years post-spray. Observations were carried out on 4 glyphosate-treated and control paired cutovers near Thunder Bay, Ontario. Moose presence and feeding activity throughout winter, as measured by periodic, systematic aerial track counts, indicated that the numbers of overwinter moose tracks were not significantly different (P > 0.05) after 0 and 1 growing seasons post-spray, but they indicated a preference for the non-sprayed control areas (P < 0.05) at 2 and 3 growing seasons post-spray. The number of moose track aggregates were similar in all control and treated cutovers (P > 0.05), prior to the first growing season post-spray, but were more numerous (P < 0.05) on control portions, 1, 2, and 3 growing seasons after treatment. Available moose browse, on control areas, was four times greater, and browse utilized was 32 times greater, than that in treated areas (P < 0.05) after 1 growing season postspray. Estimated winter moose presence calculated from pellet counts, was almost two times greater on untreated than treated areas after 1 growing season (P < 0.05) and similar at 2 growing seasons post-spray. The effect of major habitat changes brought about by forestry activities on the total moose population is discussed. It is difficult in a 3-year study to formulate conclusions based on results that take many more years to manifest themselves. Further research is recommended to determine the long-term impact of glyphosate application on wildlife habitat. Couture, G., J. Legris, and R. Langevin. 1995. valuation des impacts du glyphosate utilis dans le milieu forestier, Qubec. Gouvernement du Qubec, ministre des Ressources naturelles, Direction de lenvironment forestier, Service du suivi environmental , Qubec. See Plant and Soil Residues. RN95-3082. 187 p. Cumming, H. G. 1989. First year effects on moose browse from two silvicultural applications of glyphosate in Ontario. Alces 25: 118-32. Aerial application of glyphosate at 1.07 kg/ha on a conifer plantation released planted trees satisfactorily while reducing browse stems available to moose (Alces alces) by only 5-41%. Loss of biomass for 8 major browse species (90% of stems browsed) amounted to 96 kg/ha. Browsed stems/ha decreased from 15-82% on 5 of 6 sprayed strata, and increased by 130-3200% in 5 of 6 control strata, BUT variability rendered differences insignificant. On a second study area, ground applications of glyphosate at 2.7 kg/ha as pre-planting site preparation reduced available browse by 6392%, resulting in a loss of 145 kg/ha (for 90% of the browsed stems). Apparently, some silvicultural objectives can be achieved while retaining substantial densities of browse plants for at least the first year, but other silvicultural applications may reduce browse availability more seriously. Cumming, H. G., C. P. Kelly, R. A. Lautenschlager, and S. Thapa. 1995. Effects of conifer release with Vision (glyphosate) on moose forage quality. Alces 31 : 221-32 . During January and June, 1994, we collected twigs and leaves from 4 moose browse species growing in treated and control portions of 2 ongoing replicated block experiments in which Vision had been applied aerially at 1.60 kg a.e./ha in 1990 (4 years before our sampling), and 1.07 kg a.e./ha in 1986 (8 years before sampling). Altogether, 350 samples of forage were analyzed for crude protein and associated parameters, e.g. cutin and lignin, to calculate digestible protein and digestible dry matter. Means (and ranges) follow: crude protein - twigs 8% (7-9), leaves 15% (12-19);digestible protein - twigs 3% (2-4), leaves 8% (4-13); digestible dry matter - twigs 60% (57-63), leaves 65% (6270). Forage quality varied significantly among blocks and species, digestible protein varied between study areas in summer, but no significant differences were detected between treated and control plots either 4 or 8 years after treatment. Consistently higher values for digestible protein in summer forage from treated portions of the 8-year-old study may indicate differences that would show significance with 75

12.

13.

14.

Mammals more samples. But apart from that, the study suggests that any long-term effects of conifer release with Vision are more likely to be quantitative than qualitative. 15. Cumming, H. G., R. A. Lautenschlager, C. P. Kelly, and S. Thapa. 1996. Effects of conifer release with Vision (glyphosate) herbicide on moose forage quality (digestible protein). Ontario Forest Research Institute. Forest Research Report 139. 8 p. Digestible protein percentages in 4 important moose foods, aspen, willow, hazel, and raspberry, were not significantly reduced either 4 or 8 years after aerial treatment with silvicultural herbicide, Vision (glyphosate). Significant differences were found between seasons (winter vs. summer), study areas, and among species. Protein in summer leaves was much higher than in winter twigs. Winter values differed little between study areas, but in summer those at one study area were consistently higher than at the other, perhaps due to soil depth differences. With 1 exception, trembling aspen produced higher levels of digestible protein than any of the other 3 species examined. The lack of significant differences between treatments suggests that any nutritional changes due to applications of Vision may be short-term (less than 4 years), and then perhaps only on better sites. Therefore, longer-term (4 to 8 years post-treatment) predictions of effects of this herbicide treatment on moose browse can simply be based on browse quantities. D'Anieri, P., Jr., M. D. Leslie, and M. L. McCormack, Jr. 1987. Small mammals in glyphosatetreated clearcuts in northern Maine U.S.A. Canadian Field-Naturalist 101: 547-50. Effects of glyphosate (N-(phosphonomethyl)glycine) on small mammals in four-to-five-year-old clearcuts were evaluated by snap- and pit-trapping one area one year after treatment, one area two months before and after treatment, and one untreated control. All areas were sampled simultaneously in four trapping periods from July to October, 1984. Seven species were captured, but masked shrews (Sorex cinereus), deer mice (Peromyscus maniculatus), southern red-backed voles (Clethrionomys gapperi), and pygmy shrews (Microsorex hoyi) comprised 97% of 290 captures. Only southern redbacked voles were affected by glyphosate application, being significantly more abundant on the control and less numerous on the one-year-old spray area. No short-term changes in captures occurred after the 1984 herbicide application. Danielsen, V., and A. E. Larsen. 1990. Roundup- and Cerone-treated barley for pigs. Beretning fra Statens Husdyrbrugsforsog. No. 677. 55 p. Barley grown in 1985-1987, untreated or treated with glyphosate (Roundup), ethephon (Cerone) or both, contained glyphosate 0.8-4.0 and 9-13 mg/kg in grain and straw, respectively; ethephon was below detection limits in grain and 0.12-0.14 mg/kg in straw. In 2 experiments, 4 groups of pigs were given diets with 66-81% barley as the only cereal. In experiment 1, 21-day-old pigs were given barley diets from each of the 4 treatments for about 2.5 years, till the sows had weaned their fourth litter. In experiment 2, 48 adult sows were given the experimental diets at farrowing and for 6 months over 2 lactations and 1 pregnancy; they also had straw bedding from the corresponding treated cereals. In experiment 1, there were no glyphosate X ethephon interactions, but both increased feed intake in freely-fed weaned pigs until 10 weeks old. Feed:gain was not influenced. Feed intake and reproduction were not affected. Litter size at birth was negatively affected by ethephon, but not by glyphosate. Daily milk yield of sows was not affected. In experiment 1, there was a significant glyphosate X ethephon interaction on survival rate of piglets from sows given diets and straw bedding from barley treated with glyphosate and ethephon. Survival rate of piglets decreased from 90 to 82%. (Denmark). Eschholz, W. E., F. A. Servello, B. Griffith, K. S. Raymond, and W. B. Krohn. 1996. Winter use of glyphosate-treated clearcuts by moose in Maine. Journal of Wildlife Management 60 , no. 4 : 764-69 . Aerial treatment of naturally regenerating clearcuts with the herbicide glyphosate initially reduces the availability of deciduous browse, but may subsequently improve bedding cover for moose (Alces alces). However, the potential effects of these vegetative changes on use of clearcuts by moose has received little study. We studied effects of glyphosate treatment of clearcuts in Maine on (1) use of clearcuts by moose and (2) conifer cover during 2 periods, 1-2 and 7-11 years posttreatment. We made counts of moose tracks, beds and pellet groups on transects in treated and untreated 76

16.

17.

18.

Mammals clearcuts in January-March 1992 and 1993 and measured conifer densities in January-March 1991-93. At 1 and 2 years posttreatment, tracks of foraging moose were 57 and 75% less abundant on treated than untreated clearcuts (P = 0.013). Counts of moose beds, total tracks, and pellet groups exhibited similar patterns as tracks of foraging moose but did not differ ( P>0.1) between treatments. At 7-11 years posttreatment, tracks of foraging moose (P = 0.05) and moose beds (P=0.06) were greater on treated than untreated clearcuts. Conifer densities at 1-2 years posttreatment were not affected (P > 0.1) by treatment, but conifers 2.0-2.9 m tall were 2 times more abundant (P < 0.1) on treated than untreated clearcuts at 7-11 years posttreatment. Less foraging activity at 1-2 years posttreatment appeared to be the result of reduced browse availability because conifer cover for bedding was similar on treated and untreated clearcuts. We hypothesized that greater counts of tracks of foraging moose on older treated clearcuts was due to increased foraging activity on sites with more abundant conifer cover. 19. Evans, D. D., and M. J. Batty. 1986. Effects of high dietary concentrations of glyphosate (Roundup) on a species of bird, marsupial and rodent indigenous to Australia. Environmental Toxicology and Chemistry 5: 399-401. See Birds Section. Freedman, B. 1991. Controversy over the use of herbicides in forestry, with particular reference to glyphosate usage. Journal of Environmental Science and Health C8: 277-86. (*) In this contribution, some of the environmental effects of herbicide use in forestry are discussed, with particular reference to a commonly used organophosphonate herbicide, glyphosate. Discussion focuses on the interpretation of studies of the laboratory-based toxicology and field-based ecotoxicology of glyphosate and its silvicultural formulations. Pesticide drift and off-target deposition are notoriously variable, but they can be considerable if applications are made using inappropriate spray methods or during inappropriate meteorological conditions. Factors affecting drift and deposition are discussed. Amounts and persistence of glyphosate residues encountered after a silvicultural application are discussed and it was concluded that there is little or no propensity for glyphosate to bioaccumulate. The acute toxicity of glyphosate to mammals is similar to that of some chemicals which many humans voluntarily ingest. Effects of longer-term, chronic exposures of mammals to glyphosate are also small, especially considering doses that humans or other animals might receive during an operational treatment in forestry. Considering the relatively small acute toxicity of glyphosate to animals, it is unlikely that wildlife inhabiting sprayed clearcuts would be toxicologically affected by a silvicultural application. However, glyphosate causes great habitat changes through effects on plant productivity, and by changing the distribution of biomass in three-dimensional space and among plant species. Therefore, wildlife such as birds and mammals could be variously affected through changes in vegetation, and through secondary changes in the abundance of arthropods. These indirect effects of glyphosate spraying, which are considered within the purview of "ecotoxicology", could affect the abundance and reproductive success of wildlife on the spray site, irrespective of any direct effects due to glyphosate toxicity. The reasons for greater concern over herbicide use in forestry, considering herbicides are used to a much greater extent in agriculture, are also discussed. Gourley, M., M. Vomocil, and M. Newton. 1990. Forest weeding reduces the effect of deerbrowsing on Douglas fir. Forest Ecology and Management 36: 177-86. Three-year-old bareroot Douglas-fir [Pseudotsuga menziesii (Mirb.) Franco] transplants were established in four cutover locations in the Oregon Coast Range (USA) where deer-browsing was expected. Protection was provided against browsing by five physical treatments and one chemical treatment, each of which was tested with and without complete weed control with directed applications of glyphosate. After five years, none of the protective treatments provided any growth advantages; some even caused growth losses. In contrast, weed control, with or without additional protective measures, consistently improved growth. By the 5th year, weeded trees averaged twice the biomass of unweeded trees, regardless of browsing. Average tree size was largest in the treatment with no weed competition and with no barriers to prevent browsing. Advantages of weeding were greatest on the

20.

21.

77

Mammals poorest site. Weed control, in conjunction with the large size of transplants, appeared to prevent most loss due to damage from moderate deer-browsing. 22. Hamilton, A. N., C. A. Bryden, and C. J. Clement. 1991. Impacts of glyphosate application on grizzly bear forage production in the Coastal Western Hemlock Zone. Forest Resource Development Agreement, Forestry Canada, B.C. Ministry of Forests. FRDA Report 165. 49 p. The impacts of glyphosate on grizzly bear forage availability in coastal British Columbia were investigated. Emphasis was on floodplain ecosystems. Impacts were examined at three scales: (1) regional - occupied grizzly bear habitat in the Coastal Western Hemlock (CWH) zone; (2) local - one watershed representative of operational use; and (3) site-specific - plots in 22 cutblocks. Changes in percent cover and fruit production of bear forage species were monitored following foliar and individual stem treatments. A simple efficacy trial was also conducted. Information from the Pesticide Control Branch shows that an average of 1826 ha per year has been treated with glyphosate in occupied grizzly bear habitat in the CWH zone since 1986. Habitat mapping in the Chuckwalla-Kilbella watershed indicated that most glyphosate is used on areas of medium and high value bear forage. Foliar application, particularly by backpack or powerhose, influenced the amount of forage available for at least 3 years after treatment. Total cover of fruit-producing shrubs declined an average of 63% 1 year after operational foliar treatments. Recommendations for the integrated management of grizzly bears and forestry-related glyphosate use include: (1) using the biogeoclimatic ecosystem classification system to structure integrated management prescriptions; (2) assessing proposed treatments in the context of habitat supply within the watershed; (3) using application rates no higher than required to meet silvicultural obligations; (4) avoiding bear forage species not directly competing with crop trees; and (5) assessing alternative silvicultural methods. Integrated resource planning on floodplain site associations should consider the likelihood of silvicultural success, the economics of harvest and reforestation, and bear habitat values over time. Haugen, R., and E. Lunde. 1981. Spraying with glyphosate. I. The effect on vegetation. II. The effect on damage by small rodents. Norsk Skogbruk 27: 13-15. Results of helicopter spraying in August on plots laid out in weed-infested 7-year-old spruce plantations on various site types in Nord Trondelag, Norway are discussed. Most of the species particularly harmful to spruce were highly susceptible; a few could be classed as tolerant, including Vaccinium vitis-idaea. Grass was controlled up to 100% and herbs, including V. myrtillus, up to 70%, suggesting that treatment might deprive field mice of much of their food and (in snow-free periods) cover. Such control of vegetation developing after fellings, might reduce infestation of clear fellings by small rodents. However, it was also found that spraying increased the sugar content of spruce bark wood and needles (in the case of bark up to 41%), which might make trees more palatable to rodents. (Norway). Hjeljord, O. 1994. Moose (Alces alces) and mountain hare (Lepus timidus) use of conifer plantations following glyphosate application. Norwegian Journal of Agricultural Sciences 8 , no. 3-4 : 181-88 . Reinvasion of hardwoods and use of forest plantations by moose (Alces alces) and mountain hare (Lepus timidus) after the application of glyphosate were studied for 9 years in southwestern Norway. Hardwoods and pellets (feces) of moose and hare were recorded on permanent plots on sprayed and control sites. Nine years after spraying the number of trees was 76% of that on control sites, while shoot production remained small. Sorbus aucuparia almost disappeared on sprayed sites and on unsprayed sites was prevented from increasing in height by heavy moose browsing. Therefore, Betula sp. dominated on both sprayed and unsprayed sites 9 years after spraying. Hare use of sprayed sites decreased during the first year after spraying, but thereafter did not differ from use of control sites during the rest of the study period. Moose use of sprayed sites was lower during all years except one. (Norway) .

23.

24.

78

Mammals 25. Hjeljord, O., and S. Gronvold. 1988. Glyphosate application in forest-ecological aspects. VI. Browsing by moose (Alces alces) in relation to chemical and mechanical brush control. Scandinavian Journal of Forest Research 3: 115-21. Browse production after two growth seasons following mechanical cutting and glyphosate spraying was 60% and less than 1%, respectively, of production before treatment. Moose (A. alces) utilization of forest plantations decreased significantly the first and third winter after spraying. The second winter there was also a decrease, although not statistically significant. Browsing by A. alces relieved the spruce seedlings from considerable competition by hardwoods. Browsing also altered the hardwood composition of the forest plantations, from rowan (Sorbus aucuparia) dominance to birch (Betula sp.) dominance. (Norway). Hjeljord, O., V. Sahlgaard, V. E. Enge, M. Eggestad, and S. Gronvold. 1988. Glyphosate application in forest-ecological aspects. VII. The effect on mountain hare (Lepus timidus) use of a forest plantation. Scandinavian Journal of Forest Research 3: 123-27. The mountain hare's (Lepus timidus) use of sprayed and unsprayed parts of a forest plantation was studied using pellet counts, radiotracking and a winter feeding site survey. There was a strong reduction in use by L. timidus the first year after spraying. The second year after spraying the difference between sprayed and unsprayed parts was small by comparison. It was concluded that spraying of forest plantations up to at least 10 years after logging mainly affects the spring habitat of L. timidus. (Norway). Jones, G. B., V. L. Henderson, and K. A. Ottenbreit. 1995. Effects of aerial application of the herbicide glyphosate on forest ecosystems, part 3: Effects on vegetation and ungulate browse, 1985 to 1994. Winnipeg Manitoba Natural Resources, Forestry Branch , Manitoba Environment; Ottawa: Forestry Canada. 95 p. This report presents the results of a study initiated in 1985 to study the direct effects of aerial application of glyphosate on target and non-target vegetation in the boreal forest region of southeastern Manitoba. The investigators monitored 11 study sites 1985-94 to assess the effects of various rates of glyphosate applications on the vegetation of forest ecosystems. They established control (unsprayed) and treated sampling sites in two recently burned areas, a mature aspen stand, a forest plantation, and an area where an accidental spill of glyphosate had occurred in the past. They established a separate study to specifically address the effects of glyphosate on available ungulate browse material, collecting data on available browse from six forage species in the burn and aspen stand areas. The results presented include comparisons of plant growth in the treated and untreated areas. Jones, R., and J. M. Forbes. 1984. A note on effects of glyphosate and quinine on the palatability of hay for sheep. Animal Production 38: 301-3. Sheep were used in a Latin-square experiment to investigate the effect of preharvest treatment of pasture with glyphosate (Roundup), a translocated herbicide, on the palatability of hay. Four hays were used: control (C), Roundup (R), quinine (Q) (100 mg/kg), and Roundup with quinine (QR). During each week one of the following choices was given to each animal: (a) C/C; (b) C/R; (c) R/R; (d) Q/C; (e) Q/R; and (f) C/QR. There were no significant effects of treatment on the proportion chosen or on total daily intake, mean meal size, or rate of eating. Thus, neither preharvest treatment with Roundup nor postharvest treatment with quinine affected the palatability of hay. (United Kingdom). Jordan, P. A., E. R. Kennedy, S. D. Posner, and G. A. Weil. 1988. "In: Integrating forest management for wildlife and fish. " Integrating habitat needs of moose with timber management in northeastern Minnesota. USDA Forest Service. General Technical Report NC122. pp. 18-22. Where softwood production is being emphasized, plantations are established, and competing hardwoods are suppressed with herbicides. These hardwoods, however, are critical forage for moose, an important wildlife resource of north-eastern Minnesota. Moose favor young plantations for the high density of forage there. It was shown that spraying 2,4-D reduces browse availability for several years by about half, and that the recently more popular chemical, glyphosate, reduces browse by some threequarters and for a longer span. Studies of growth in white spruce relative to varying levels of surrounding shrubs suggest that current Forest Service criteria for herbicide applications are more strict 79

26.

27.

28.

29.

Mammals than necessary. In Minnesota emphasis is now shifting from mainly softwood production to more use of aspen; its regeneration requires no plantations or herbicides. It is suggested that a timber system optimally beneficial to moose should include mixed-species interspersions. Application of new techniques such as computerized mapping coupled with modelling will be tested for achieving better integrated resource management. 30. Kammerer, M. 1995. Intoxication by glyphosate based herbicides. Recueil De Medecine Veterinaire De L'Ecole D'Alfort 171: 149-52. See Human Health Section. Kelly, C. P. 1993. "Effects of variable rate aerial application of Vision on moose (Alces alces) browse and conifer crop tree performance. " M.Sc.F. Thesis, Lakehead University. 102 p. Experimental aerial treatment of 7 mixedwood areas in late summer for conifer release with Vision at 0.80, 1.06, and 1.60 kg a.e./ha, decreased living hardwood stem densities after one winter by 42, 61 and 42% respectively on treated plots, while controls increased by 13%. Two winters after treatment stem densities were reduced (from pre-spray levels) by 48, 65 and 61%; controls increased 19%. Greatest numbers of stems occurred on moderately deep, fresh soils. Winter browsing rates decreased on all plots after treatment and were consistently higher on controls when compared with treated sub-blocks. Decline was progressive over two years after treatment on sprayed areas but recovered in the second year on controls. The two highest application rates had the lowest browsing levels. Conversely, winter track data showed no differences in moose use between sprayed areas and controls, nor any difference among treatments. This suggested moose still travelled through sprayed areas, but did not stop to browse. In addition to stem density counts, cover (%) for both herbs and hardwoods were estimated to evaluate the effectiveness of Vision as a conifer release. Hardwood cover was reduced significantly by all application rates; differences among treatments were not significant. Herbaceous ground cover was reduced approximately 20% on all treated areas one season after spray but by next year these sprayed areas had recovered to equivalent levels as controls. Neither crop tree diameter nor height growth was affected by Vision application at this early stage in the experiment. Moose densities within these study areas appear to be low enough that food is not a limiting factor. The small amount of spraying in Ontario (relative to the productive forest land base) is not expected to impact moose populations. However, in areas with high concentrations of sprayed cutovers there should be concern. Results of this short term study suggest that 0.80 kg a.e./ha controlled hardwood and herbaceous competition as well as 1.06 & 1.60 kg a.e./ha. However, the lowest application rate showed signs of increased moose use two years post spray compared with the two higher rates. Consequently, when spray programs are concentrated in one management unit, it is recommended to spray at 0.80 kg a.e./ha. Kelly, C. P., and H. G. Cumming. 1992. Effects of an aerial application of Vision on moose browse - first year results. Alces 28: 101-10. Experimental aerial treatment of 7 mixedwood areas for conifer release with Vision at 0.80, 1.06, and 1.60 kg a.e./ha, decreased living stem densities after one winter by 36, 61 and 47% respectively on treated plots, while controls increased by 25%; thus, decreases in total numbers of living hardwood stems were not proportional to application rates. Hardwood shrub cover was reduced by application rates of 1.06 and 1.60 kg a.e./ha, but not by 0.80 kg a.e./ha. Although 1.06 kg a.e./ha reduced cover almost twice as much as the highest concentration, differences were not significant. Herbaceous ground cover was reduced approximately 20% on all treated areas when compared with control plots. Browsing rates decreased on all plots after treatment and were twice as high on controls when compared with treated sub-blocks. However, neither differences among treatments, nor between treated and control plots were significant. Kennedy, E. R., and P. A. Jordan. 1985. Glyphosate and 2,4-D: The impact of two herbicides on moose browse in forest plantations. Alces 21: 149-60. Conifer plantations in northeastern Minnesota are important browse areas for moose. The U.S. Forest Service has recently shifted to glyphosate (Roundup) as the predominant herbicide for controlling hardwood shrub competition in plantations. Glyphosate is a systemic toxin that kills the entire plant so little resprouting occurs. Previously, the preferred herbicide

31.

32.

33.

80

Mammals was 2,4-D, which allows vigorous sprouting. We sampled available browse in glyphosate and 2,4-D treated stands. Three years after spraying, the glyphosate stands averaged only half the available browse as the 2,4-D stands. While sensitivity to 2,4-D differs markedly among woody plant species, glyphosate kills woody species more uniformly. Grass and raspberries are not controlled one year after spraying because glyphosate has no residual effects. We could not measure longterm effects in this study because glyphosate was not used in this region before 1981. This report covers only the first half of a 2-year study. 34. Lautenschlager, R. A. 1993. Effects of conifer release with herbicides on wildlife. (A review with an emphasis on Ontarios forests). VMAP Forest Research Information Paper No.111 . Ontario Ministry of Natural Resources, Sault Ste. Marie, Ontario. See Biodiversity and Habitat Restoration Section. . 1986. "In: Is Good Forestry Good Wildlife Management? J. Bissonette (ed.) ." Forestry, herbicides, and wildlife. Maine Agriculture Experimental Station, University of Maine, Orono. Miscellaneous Publication No. 689. pp. 299-307. At present application rates, herbicide conifer release in the northeast probably affects wildlife populations little. However, if herbicide conifer release increases greatly, it will: reduce the ability of treated areas to support moose (Alces alces) and small mammal populations for a short period following treatment; at most cause a short term reduction of deer (Odocoileus virginianus) and hare (Lepus americanus) populations, which are more likely to benefit within a few years after application; and benefit warblers and associated bird species normally found in the spruce/fir (Picea spp./Abies balsamea) forest. Although herbicide conifer release could benefit a variety of forest wildlife, forestry herbicides have developed an unjustified negative reputation. That reputation is based on: environmental fears and generalizations developed in the 1960's; herbicides' incorrect association with other, more toxic, chemicals (e.g., insecticides); and the inability of some resource workers and the media (and therefore the public) to distinguish among the variety of pesticides. To help eliminate this confusion it is imperative that resource professionals use "pesticide" only when a more specific term (insecticide, herbicide, fungicide etc.) is inappropriate. . 1991. Response of wildlife in northern ecosystems to conifer release with herbicides. Cooperative Forestry Research Unit , University of Maine, Orono, Maine. Information Report 26. 12 p. See Birds Section. . 1992. Effects of conifer release with herbicides on moose: browse production, habitat use, and residues in meat. Alces 28: 215-22. Six studies, 5 in spruce plantations and 1 in naturally regenerated spruce-fir stand, have examined the effects of conifer release with herbicides on moose browse production and habitat use. Both were reduced in plantations and naturally regenerated spruce-fir stands for up to 4 growing seasons after treatment. Only 1 study, in a naturally regenerated stand, examined long-term effects, and there forage production on all treated areas exceeded production on controls 8 growing seasons after treatment. Although feeding studies and residues in digestive tracts show that animals consume some glyphosate while feeding, herbicides were not found in the flesh of game animals (moose, deer, hare) taken from within or near areas released with glyphosate. . 1993. Response of wildlife to forest herbicide applications in northern coniferous ecosystems. Canadian Journal of Forest Research 23: 2286-99. Reviewed studies of the effects of forest herbicide applications on wildlife often lacked replication, pretreatment information, and (or) were conducted for only one or two growing seasons after treatment. Because of these problems, as well as the use of dissimilar sampling techniques, study conclusions have sometimes been contradictory. A review of eight studies of the effects of herbicide treatments on northern songbird populations in regenerating clearcuts indicates that total songbird populations are seldom reduced during the growing season after treatment. Densities of species that use early successional brushy, deciduous cover are sometimes reduced, while densities of species which commonly use more open areas, sometimes increase. A review of 14 studies of the 81

35.

36.

37.

38.

Mammals effects of herbicide treatments on small mammals indicates that like songbirds, small mammal responses are species specific. Some species are unaffected, while some select and others avoid herbicide-treated areas. Only studies that use kill or removal trapping to study small mammal responses show density reductions associated with herbicide treatment. It seems that some small mammal species may be reluctant to venture into disturbed areas, although residents in those areas are apparently not affected by the disturbance. Fourteen relevant studies examined the effects of conifer release treatments on moose and deer foods and habitat use. Conifer release treatments reduce the availability of moose browse for as long as four growing seasons after treatment. The degree of reduction during the growing season after treatment varies with the herbicide and rate used. Deer use of treated areas remains unchanged or increases during the first growing season after treatment. Eight years after treating a naturally regenerated spruce-fir stand browse was three to seven times more abundant on treated than on control plots (depending on the chemical and rate used). Forage quality (nitrogen, ash, and moisture) of crop trees increased one growing season after the soilactive herbicide simazine was applied to control competition around outplanted 3-year-old balsam fir seedlings. 39. . 1997. Effects of alternative conifer release treatments on small mammals in northwestern Ontario. Forestry Chronicle 73 (In press) . Density changes were quantified in populations of small mammals responding to different conifer release treatments [manual (brushsaw) cutting; mechanical (Silvana selective) cutting; aerialapplied herbicides (Release (a.i. triclopyr), Vision (a.i. glyphosate)); and no treatment (control)]. Nearly 4,000 individual small mammals were captured and released during the first two years of this study. The most commonly captured were deer mice (Peromyscus maniculatus), redback voles (Clethrionomys gapperi), and common shrews (Sorex cinereus). Short-tailed (Blarina brevicauda), black-backed (S. arcticus) and pygmy (S. hoyi) shrews; eastern (Tamias striatus) and least (T. minimus) chipmunks; meadow voles (Microtus pennsylvanicus); short-tailed weasels (Mustela erminea); and meadow jumping mice (Zapus hudsonius) were fairly common. During the first growing season after treatment deer mouse densities were highest on Release, Silvana, and brushsaw plots, and lower on control and Vision plots. Redback vole densities were highest and very similar on control, brushsaw, and Silvana plots, intermediate on Vision plots, and lowest on the Release-treated plots. At this time eastern chipmunk densities were highest on Vision and control plots, intermediate on brushsaw and Silvana plots, and lowest on Release-treated plots. One growing season post-treatment, small mammal responses to proposed alternative conifer release treatments varied somewhat, but were similar to responses to the standard (Vision herbicide) conifer release treatment. Lautenschlager, R. A., C. Hollstedt, and F. W. Bell. 1995. Effects of herbicide, manual, and annual release of young jack pine on vegetation and small mammals in northwestern Ontario. Second International Conference on Forest Vegetation Management, Rotorua, New Zealand. R.E. Gaskin, J.A. Zabkiewicz (compilers), FRI Bulletin No. 192. pp. 149-50. This study documents the effects of standard aerial herbicide (glyphosate) application, annually repeated herbicide applications and a proposed alternative (brushsaw cutting) for conifer release in jack pine systems in northwestern Ontario. Treatments increased survival and growth of planted pine, but their effects on resident small mammals were species specific. Redback voles and eastern chipmunks were reduced, deer mice were unaffected, and least chipmunks increased following herbicide treatments. Manual brushsaw cutting did not reduce small mammal populations less, or for a shorter period, than traditional aerial release treatments. Legris, J., and G. Couture. 1991. Rsidus de glyphosate dans le gibier (livre, orignal et cerf de Virginie) suite des pulvrisations en milieu forestier en 1988. Ministre des Forts, Service des analyses environnementales, Gouvernement du Qubec. 91-3016. 24 p. In 1988 the forests sector of the ministre de l'nergie et des Resources looked into the possible contamination of game animals resulting from the use of glyphosate in softwood plantation maintenance operations in public forests. The glyphosate was applied at 1.5 kg AI/ha. Samples of snowshoe hare (Lepus americanus), moose (Alces alces) and white-tailed deer (Odocoileus virginianus), shot inside or close to the treated areas, were collected in the vicinity of Rimouski and Matane. Sampling was done in October, during hunting season and approximately two 82

40.

41.

Mammals months after the treatments had been applied (in August). Flesh, liver, kidney, urine, stomach content and feces samples were removed from 19 hares. Flesh, liver and kidney samples were taken from 16 dead moose. Only one white-tailed deer flesh sample was obtained. For the three species under study, 31 of the 32 flesh samples showed no detectable residue. The detection threshold for these analysis was 0.050 g/g (wet weight). The only positive value (0.146 g/g) was found in a moose flesh sample. However, this reading may be the result of accidental contamination during the sampling procedure and may have been caused by the presence of moose hairs on the sample. Stomach content samples were taken from 17 hares. Six of these showed glyphosate residues ranging from 0.084 to 0.262 g/g. These concentrations represent less than 20% of the glyphosate that may be found in vegetation approximately two months after treatment. Analysis of the 18 liver samples revealed no trace of residue. Analysis of the 19 kidney samples revealed only one positive value (0.208 g/g). Two of the seven urine samples contained glyphosate (109 and 142 g/L). Residues ranging from 0.174 to 3.52 g/g were found in 13 of the 15 feces samples. None of the 16 moose liver samples or nine kidney samples showed any detectable residue. The results therefore indicate that glyphosate can be ingested through treated vegetation and is mainly eliminated through the urinary and fecal tracts. In general, these results concur with the observations reported in the literature. Given the product's mode of action, its characteristics, and the low residue levels found in the samples, we do not anticipate any particular impact on the species studied. The generalized absence of detectable residues in the flesh, liver and kidneys seems to confirm that the risk of contamination from the consumption of these meats is very low. 42. Lveill, P., J. Legris, and L. Deschenes. 1996. Exploratory study of glyphosate residues in small mammals after an aerial forestry application. B-102. Gouvernement du Qubec, Ministre des Ressources naturelles, Qubec. In general, species used in this study turned out to be good indicators of the presence of residues shortly after treatment. The deer mouse (Peromyscus maniculatus), the red-backed vole (Clethrionomys gapperi) and the gray shrew (Sorex cinereus) did indicate the presence of residues. The grand shrew (Blarina brevicauda) never did show any detectable residues. Residue levels were of the order of 1 ppm or less, 15 days after treatment. At 45 days after treatment, residues were detected less frequently. Residue levels found are similar to those previously reported in the literature. Based on available information, these residue levels do not constitute a risk of either acute or subchronic toxic effects for these small mammals, nor for their predators. Lloyd, R. A. 1989. Assessing the impact of glyphosate and liquid hexazinone on moose browse species in the Skeena Region. B.C., Third year report. Ministry of Environment, Fish and Wildlife Branch, Smithers, B.C.. This report is presented in three parts. Part A shows how herbicide impact may be described as the range of effects produced on a given species in a given area. The range of effects possible on one individual in the first year following treatment were Light, Moderate, Severe, Very Severe or Dead. The term "damage assessment" was used to portray the percentage of individuals of a given species in each category. In general, Light, Moderate and some Severe individuals were found to recover, whereas Dead, Very Severe and some Severe individuals did not. Thus, when a site is assessed by this method, it is possible to show not only the range of damage, but also the expected degree of recovery. Part B describes how moose utilization of an area depends on herbicide impact. Moose were found to browse Light or Moderate plants, not Severe, Very Severe or Dead. Thus, where most browse plants were Light or Moderate, the area will continue to be useful to moose; where most were Severe, Very Severe or Dead, much of its value is lost. Herbicide residues in treated browse were up to 73 g/g 4 weeks after treatment, and undetectable the following year. Toxic effects on moose seem unlikely Part B also shows that willow and red-osier dogwood are comparatively resistant to glyphosate, and may generally be expected to show 25-60% recovery when sprayed at 1.8-2.1 kg a.i./ha. Aspen, cottonwood, maple and birch are much less resistant and may show over 90% mortality under the 83

43.

Mammals same conditions. Invasions of herbaceous pioneer species sometimes occurred in the second year after treatment; this seemed typical of moist sites which suffered heavy impact. This suggests that application rates of 1.8-2.1 kg a.i./ha are too high and are causing unnecessarily severe damage to non-target browse and non-browse plants. Part C makes some recommendations as to how impact on browse species may be minimized. Most of these involve using a lower concentration, and not treating entire areas at once. 44. . 1990. Assessing the impact of glyphosate and liquid hexazinone on moose browse species in the Skeena region. Addendum., B.C. Ministry of Environment, Fish and Wildlife Branch, Smithers, B.C.. The impact of herbicide treatment on moose habitat can be approached from two different directions; namely, changes in browse availability and composition which are directly attributable to herbicide use, and changes in the way that moose utilize a treated area. Damage assessments demonstrate herbicide impact on shrub species, especially browse species, in the year following treatment. Some further mortality of severely impacted plants can be expected in the second year and, at least for willows and Douglas maple, in the third year as well. At Smithers Landing, 3 1/2 years after treatment, moose utilization as determined from track counts was approximately eight times higher in the control area than in the treated. There are approximately twice as many live willow stems in the control area as in the treated, although the areas were probably similar prior to treatment. It seems that herbicide use did diminish the value of the treated area as moose habitat, and that the decrease was not solely correlated with browse abundance. In future, herbicide impact on moose habitat should also be considered in terms of habitat selection and not only in terms of browse. Out of fifteen sites first surveyed in 1987 and 1988, eleven showed vigorous herbaceous growth at least in some parts of the cutblock. Herbaceous growth was most obvious in moist areas where most of the shrub vegetation existing prior to treatment had been killed. The commonest species were fireweed and grasses, especially Elymus glaucus, Calamagrostis canadensis, Festuca spp. and Bromus vulgaris. This could give rise to concern from both forestry and wildlife management perspectives. Shrub establishment from seed was not observed at any of the sites visited. . 1990. Impact on vegetation after operational Vision treatment at varying rates in the Skeena region. B.C. Ministry of Environment, Fish and Wildlife Branch, Smithers, B.C. Four cutblocks each received Vision treatments at various application rates in August 1988. The Blunt Fire, an established mixed brush community to 6 m tall, received 2, 3, 4 and 4.5 litres of formulated product per hectare; Taltapin Lake, an established mixed brush community on a dry site, received 3, 4 and 5 l/ha; Tachek Creek, a rehab site completing its second growing season after mechanical disturbance and broadcast burn, received 3 and 5 l/ha; and Pinkut Creek, an established mixed brush community on a dry-to-moist hillside site received 3 and 5 l/ha. A 4 l/ha treatment was also applied to Pinkut Creek but could not be clearly distinguished in the field. Data have been gathered before treatment and one year after treatment. Damage sustained seemed to depend on site conditions as well as on application rate. Vegetation and site conditions were rarely uniform even within a site, especially at Taltapin Lake and Pinkut Creek. At the Blunt Fire, little impact was seen after most treatments on willow and red-osier dogwood; considerable impact was seen on aspen, cottonwood, Sitka alder and fireweed; and very high impact was seen on birch and thimbleberry. The 4.5 l/ha and 2 l/ha rates caused the most impact; the 3 l/ha and 4 l/ha caused least. At Taltapin Lake, none of the treatments caused extensive damage. Some plants were showing signs of stress prior to treatment. Little impact was seen on willow of Sitka alder after treatment at 3 l/ha or 5 l/ha; aspen showed more impact, especially at 4 l/ha. Fireweed showed 35%40% reduction in percent cover at 3 l/ha and 5 l/ha. Vegetation was comparatively tall and dense in the 3 l/ha, and comparatively sparse in the 4 l/ha unit. The 4 l/ha caused the most impact and the 3 l/ha caused the least At Tachek Creek, considerable impact was seen on all species, especially black twinberry, aspen and Sitka alder. Impact was somewhat less on willow and red-osier dogwood. Little difference

45.

84

Mammals was detected between treatment at 3 l/ha and 5 l/ha. The herb layer, consisting mostly of grasses, fireweed and Watson's willowherb, was growing vigorously in both areas. At Pinkut Creek, birch showed very high impact at both rates. Sitka alder and aspen also showed high impact. Red-osier dogwood showed less impact, especially where it was growing underneath an aspen canopy. Little difference was detected between the two rates, although this may have been because the high rate was applied in a dry area and the low rate was applied in a moister area. Browse assessments at the Blunt Fire showed no clear changes in biomass of new willow shoots produced before and after treatment, although shoots produced after treatment were smaller and more numerous, especially in the 4.5 l/ha unit. Browsing was much less in the winter after treatment than in the winter before, and could have contributed to the increased number of smaller shoots. Vision treatments could be made at well below the legal maximum rate, depending on wildlife and silvicultural objectives, vegetation condition and site conditions. Low application rates should be considered: (1) where the major target species is susceptible to Vision (2) where high impact is not required (as in an established plantation) (3) on moist sites where impact is expected to be higher than average (4) where target vegetation is growing vigorously at the time of treatment. 46. Marrs, T. C. 1993. Organophosphate poisoning. Pharmacology & Therapeutics 58 , no. 1 : 5166. See Human Health Section. McComb, W. C., L. R. Curtis, K. Bentson, M. Newton, and C. L. Chambers. 1997. Toxicity of glyphosate herbicide to terrestrial mammals and amphibians of the Oregon Coast Range. (Submitted) . See Biodiversity and Habitat Restoration Section. McComb, W., L. Curtis, K. Bentson, M. Newton, and C. Chambers. 1990. Toxicity analyses of glyphosate herbicide on terrestrial vertebrates of the Oregon coast range., U.S.D.A. Forest Service, Oregon State University. 30 p. (*) Except for treated newts using dead wood cover more frequently than control animals and treated chipmunks using more early seral stage habitat than control chipmunks, there was no detectable difference in field behavior or survival between Townsend's chipmunks or rough-skinned newts receiving a high sub-lethal dose of glyphosate and those receiving a control substance. Given the high degree of consistency in direct toxic effects among the species we examined, we suspect that few, if any, sub-lethal effects would be detected in the other species that we studied in the direct toxicity effects portion of this study. There may be effects on other aspects of the field biology of these animals, such as reproductive rates. But recent field data indicate that changes in habitat quality following herbicide application can result in high reproductive activity in species associated with the grasses and forbs that proliferate following field applications (Sullivan 1990, Figures 1-4). Indeed, habitat for deer mice, Oregon voles, and Townsend's chipmunks may be enhanced following clearcutting. Oregon vole habitat may be further enhanced by subsequent spraying with glyphosate (Sullivan 1990). Although direct and indirect effects of exposure of terrestrial vertebrates to glyphosate are possible during forestry applications, we were unable to detect potential adverse effects on 7 of 9 wild species. Data for 2 species (tailed frogs and Pacific giant salamanders) were too few to allow predictions of direct effects; however the LD50 for tailed frogs seems to be higher than that of the other amphibian species. McMillan, L. M., J. F. Connor, H. R. Timmermann, J. G. McNicol, and C. S. Krishka. 1990. Small mammal and lesser vegetation response to glyphosate tending in North Central Ontario. Ontario Ministry of Natural Resources, Northwestern Ontario Forest Technology Development Unit, Ontario. Technical Report No. 57. 28 p. Small mammal response to herbicide-induced habitat changes was studied in four northcentral Ontario clearcuts. Eight hundred and sixty-seven small mammals representing eight species were sampled in 3,600 trap nights over three years. Deer mice (Peromyscus maniculatus) accounted for

47.

48.

49.

85

Mammals 55% of all captures followed by least chipmunks (Eutamias minimus) 16%, red backed voles (Clethrionomys gapperi) 15%, three species of shrews 10%, and meadow voles (Microtus pennsylvanicus) and jumping mice (Zapus hudsonius) 4%, collectively. More specimens were captured on glyphosate-treated clearcuts, one, two, and three years post-treatment than on untreated controls (475 versus 392). When plots located in missed strips and residual vegetation within sprayed clearcuts were removed from the dataset, the treatment effects appeared minimal (404 treated versus 387 control). Deer mice increased two and three years post-spray on the treated plots, whereas red-backed voles tended to decrease one, two, and three years post-spray. It is postulated that these variations were related to changes in vegetation due to herbicide treatment. Seventy-one species/genera of woody and herbaceous vegetation were assessed over three years of the study. Species richness was variable, but there were no significant changes over time. The relationship between small mammals and vegetation was discussed. 50. Miller, K. V., and J. S. Witt. 1991. Impacts of forestry herbicides on wildlife. Proceedings of the Sixth Biennial Southern Silvicultural Research Conference, S.S. Coleman and D.G. Neary (editors). Asheville, North Carolina : Southeastern Forest Experiment Station. Vol. 2. pp. 795800. See Biodiversity and Habitat Restoration Section. Milton, G. R., and J. Towers. 1990. Relationships of songbirds and small mammals to habitat features on plantation and natural regeneration sites. Canadian Institute of Forestry, St. Mary's River Forestry-Wildlife Project. Report No. 7. 57 p. (*) There are readily identifiable differences between our natural regeneration (RGN) and plantation (PLN) sites which may be attributed to the effects of herbicide (glyphosate) application. In general, the duration broad-leaved woody shrubs dominate the overstory of a site is reduced on PLN following herbicide application. There is a pronounced difference in the vertical, horizontal, and ground cover profiles compared to RGN which continues for a number of years afterwards. Silvicultural herbicide application reduces their density and percent cover but does not eliminate broad-leaved woody shrubs from PLN sites. Many plants do not succumb and reinitiate growth the following season, others seed in, and swaths of vegetation can be missed by the spray. While there are changes in the percent composition of non-woody ground cover, there is no reduction in the heterogeneity of cover types between conditions. The greatest effect on the bird community occurs when the mature forest is cut, and a major determinant in subsequent bird communities at different seral stages is the density of cull and standing dead trees We do however predict a divergence in the composition and relative abundance of species on our RGN and PLN sites during the intermediate stages of different successional paths to a conifer dominant mature forest. Species composition of small mammals on RGN and PLN sites were the same and do not differ with other studies on either mature and naturally regenerating conifer forests or plantations. Total and individual species relative abundance is more variable and is probably a response to changes in microhabitat features. Most species of small mammals occur under a wide range of habitat conditions and even large scale disturbances (such as clearcutting) will have only a temporary impact on species composition. Moreover, the relative abundance of individual species will vary with the type of ground cover which can be influenced by forest management practices. Morrison, M. L., and E. C. Meslow. 1983. Impacts of forest herbicides on wildlife: Toxicity and habitat alteration. Transactions of the 48th North American Wildlife Conference. pp. 175-85. Our review indicates that the response of wildlife to herbicide-induced habitat change is extremely varied. Variations in response by wildlife are understandable given that plants respond in a species-specific manner to the chemical applied, the rate and time of application, and various environmental constraints. What generalizations can be drawn then? First, we have seen that certain animals respond to habitat alteration by increasing their use of undamaged vegetation. Other species, however, are seemingly unable to compensate for habitat loss and thus decline in density. Other species respond to habitat change by increasing in density. This shift in concentration of density is due, quite simply, to changes in the amount of preferred habitat available to each species following treatment. An increase in density by certain species following treatment is not necessarily desired, for 86

51.

52.

Mammals this change has been artificially induced by a management practice. Increasing deer browse through herbicide application, for example, would also cause a decreased habitat availability for other species. The general response of wildlife to herbicide application can thus be predicted if data are available on the range of habitats occupied by a species and their density in these habitats. Changes anticipated in animal communities can be alleviated, in part, by careful planning of treatment. What is generally deemed desirable is retention of the natural variety of vegetation types so that managed lands can supply a diversity of vegetation and wildlife through time. 53. Newton, M., E. C. Cole, R. A. Lautenschlager, D. E. White, and M. L. McCormack, Jr. 1989. Browse availability after conifer release in Maine's spruce-fir forests. Journal of Wildlife Management 53: 643-49. Even-aged management in northern spruce (Picea spp.)-fir (Abies spp.) forest types may result in significant production of woody vegetation < 2.5 m tall (i.e., available to browsers) for > 10 years after cutting. Rapid growth of hardwood and associated conifers often leads to good browse conditions 7-9 years after harvest, followed by rapid reduction of available browse biomass (foliage and twigs) by the sixteenth year. August application of glyphosate or triclopyr and/or phenoxy herbicides 7 years after winter "clearcut" harvest reduced the cover of tall hardwoods and caused a 4- to 8-fold increase in available browse. Nearly all plant species present before herbicide treatment remained > 9 years later, but browse available after treatment included a higher proportion of sun-grown foliage and twigs than was found in untreated control plots. Spacing in conifer stands by planting or precommercial thinning should prolong the period of browse availability to > 20 years. Our data indicate that intensive management designed to increase forest crop production in northeastern spruce-fir ecosystems also dramatically improves conditions for resident browsers. Newton, M., K. M. Howard, B. R. Kelpsas, R. Danhaus, C. M. Lottman, and S. Dubelman. 1984. Fate of glyphosate in an Oregon USA forest ecosystem. Journal of Agricultural and Food Chemistry 32: 1144-51. Glyphosate herbicide residues and metabolites were evaluated in forest brush field ecosystems in the Oregon coast range aerially treated with 3.3 kg/ha glyphosate. Deposits were recorded at various canopy depths to determine interception and residues in foliage, litter, soil, streamwater, sediments and wildlife for the following 55 days. The half-life of glyphosate ranged from 10.4 to 26.6 days in foliage and litter and twice as long in soil. The treated stream peaked at 0.27 mg/L and decreased rapidly; concentrations were higher in sediment than in water and persisted longer. Coho salmon fingerlings did not accumulate detectable amounts. Exposure of mammalian herbivores, carnivores and omnivores and retention of herbicide seemed to vary with food preference; however, all species had visceral and body contents at or below observed levels in ground cover and litter, indicating that glyphosate will not accumulate in higher trophic levels. (Aminomethyl)phosphonic acid was found at low concentrations but degraded rapidly. N-nitrosoglyphosate was nondetectable. Nickerson, N. H. 1992. Impacts of vegetation management techniques on wetlands in utility rights-of-way in Massachusetts. Journal of Arboriculture 18: 102-6. See Biodiversity and Habitat Restoration Section. Pojar, R. 1990. The effects of operational application of the herbicide glyphosate on target brush species, selected browse species and conifer crop trees. 3. Two years after application. Ministry of Forests Progress Report, B.C. 49 p. (*) The results of this study show that the response of browse plants to glyphosate application varies considerably from block to block. It is impossible to draw conclusions from studies such as this as to what are the primary factors influencing the responses observed. There are just too many variables. A firmer handle on what the important factors are can probably only be obtained by carrying out more carefully controlled experiments including rate and timing trails on individual species. Because of this variability, results obtained from studies on operational applications can only be used with extreme caution to predict the response of browse to glyphosate applications.

54.

55.

56.

87

Mammals 57. Raymond, K. S., F. A. Servello, B. Griffith, and W. E. Eschholz. 1996. Winter foraging ecology of moose on glyphosate-treated clearcuts in Maine. Journal of Wildlife Management 60 , no. 4 : 753-63 . The herbicide glyphosate is widely used in northern coniferous forests of the United States and Canada to promote conifer dominance on clearcut sites by suppressing regeneration of deciduous species. We determined effects of glyphosate treatment of regenerating clearcuts on (1) browse availability (total biomass, by species, and proportion with high digestible energy [DE] content), (2) browse use, and (3) diet quality of moose (Alces alces) in winter in Maine during 2 periods: 1-2 and 711 years posttreatment. We measured browse availability and use and collected browse samples for nutritional analyses on 12 clearcuts in January-March 1991 before aerial treatment with glyphosate of 6 of these clearcuts in August 1991. We conducted posttreatment sampling of treated and untreated clearcuts during January-March 1992 and 1993. We also sampled 14 clearcuts that had been treated with glyphosate 7-11 years earlier and 5 untreated clearcuts of similar age in January-March 1992 or 1993. Available biomass (kg/ha) of deciduous browse decreased (P = 0.001) 70% on treated clearcuts relative to untreated clearcuts from pretreatment to year 2, but was not affected (P = 0.29) at 7-11 years posttreatment. Available browse from red maple (Acer rubrum) and paper birch (Betula papyrifera) appeared to decrease less than pin cherry (Prunus pensylvanica) in years 1-2 suggesting that species composition on sites may influence the magnitude of effects on total browse availability. The proportion of deciduous browse biomass with a relatively high DE content (1.8 kcal/g) was not affected (P = 0.37) by treatment at 1-2 years, but was greater (P = 0.047) on treated than untreated clearcuts at 7-11 years posttreatment. Biomass and percent of available deciduous browse eaten by moose were not affected (P > 0.1) by glyphosate in years 1-2, but were 4-5 times greater (P < 0.1) on treated than untreated clearcuts at 7-11 years posttreatment. The DE and protein content of moose diets on clearcuts was not affected (P > 0.1) by treatment in either time period. Initial reductions in browse availability may decrease the suitability of clearcuts for foraging by moose, but this effect would decrease over the next 5-9 years because browse availability decreases naturally on untreated sites. We concluded that glyphosate did not have important effects on diet quality. Heavy browsing in older treated clearcuts suggests that moose may be attracted to these sites, but this behavior was not directly related to browse availability or nutrition. We discuss management options for minimizing effects of glyphosate treatment on moose habitat. Ritchie, D. C., A. S. Harested, and R. Archibald. 1987. Glyphosate treatment and deer mice in clearcut and forest. Northwest Science 61: 199-202. Prior to planting conifers, herbicides are commonly used to reduce competition from deciduous trees and shrubs. Herbicides are usually not toxic to wildlife but do affect their habitats. We examined deer mice (Peromyscus maniculatus) to assess the impact of herbicides on small mammals. Deer mice from adjacent untreated and glyphosate-treated clearcuts had similar body sizes and numbers of placental scars and foeti. In untreated clearcut, deer mice were more abundant than in treated clearcut, but were less abundant than in surrounding old growth forest. Glyphosate altered vegetation and reduced density of deer mice in young seral stages. Habitat changes induced by glyphosate likely modified abundance and quality of food and cover for small mammals. Roy, D. N., S. K. Konar, S. Banerjee, D. A. Charles, D. G. Thompson, and R. Prasad. 1989. Uptake and persistence of the herbicide glyphosate (VisionR) in fruit of wild blueberry and red raspberry. Canadian Journal of Forest Research 19: 842-47. See Plant and Soil Residues Section. Runciman, J. B., and T. P. Sullivan. 1996. Influence of alternative conifer release treatments on habitat structure and small mammal populations in south central British Columbia. Canadian Journal of Forest Research 26, no. 11: 2023-34. This study was designed to test the hypothesis that conifer release treatments would simplify habitat structure and reduce small mammal populations in forest plantations. A secondary objective was to examine some important demographic characteristics, for selected small mammal species, that may be affected by changes in habitat. We examined the effects of manual cutting and cut-stump applications of glyphosate herbicide on vegetation, woody debris, and small mammal populations from 1991 to 1994 in young mixed-conifer plantations of south central British Columbia, Canada. The 88

58.

59.

60.

Mammals experimental design consisted of 9 separate and independent plantations: 3 controls, 3 manual treatments, and 3 cut-stump treatments. Total volumes of herbs, shrubs, coniferous trees, and woody debris were not affected by manual or cut-stump treatments for conifer release. Both treatments reduced total volumes of deciduous trees in the first posttreatment year. However, deciduous tree volumes on manual treatments had largely returned to pretreatment levels by the second posttreatment year. There were no significant (P>0.05) effects of manual or cut-stump treatments on the population size of deer mice (Peromyscus maniculatus Wagner), yellow-pine chipmunks (Tamias amoenus J.A. Allen), southern red-backed voles (Clethrionomys gapperi Vigors), or long-tailed voles (Microtus longicaudus Merriam). The response of meadow voles (Microtus pennsylvanicus Ord) was variable. Sex ratios, body weights, reproduction, recruitment, and survival of deer mice remained similar on treatment and control plantations throughout this study. Changes in habitat structure up to 2 years posttreatment did not appear to exceed the tolerance of small mammal populations for early successional change. 61. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. Wageningen, NL: European Weed Res. Soc. pp. 315-21. See Water Quality Section. Santillo, D. J., D. M. Leslie, Jr. , and P. W. Brown. 1989. Responses of small mammals and habitat to glyphosate application on clearcuts. Journal of Wildlife Management 53: 164-72. We investigated effects of herbicide-induced habitat changes on small mammals in clearcuts in northcentral Maine. Fewer small mammals were captured on glyphosate (nitrogen-phosphonomethyl glycine) (Roundup, Monsanto, St. Louis, Mo.)-treated clearcuts 1-3 years post-treatment compared to untreated clearcuts. Insectivores (Soricidae) comprised 72% of small mammal captures and were less abundant (P < 0.001) for all 3 years post-treatment. Herbivores (Microtinae) were less abundant 1 (P < 0.01) and 2 years (P < 0.001) post-treatment. Omnivores (Cricetinae and Zapodidae) were equally abundant on treated and untreated clearcuts. Differences in small mammal abundance paralleled herbicide-induced reductions in invertebrates and plant food and cover. Patches of untreated vegetation within herbicide-treated clearcuts provided a source of invertebrates and plant food and cover. Santillo, D.J. 1994. Observations on moose, Alces alces, habitat and use on herbicide-treated clearcuts in Maine. Canadian Field-Naturalist 108 , no. 1 : 22-25 . I studied Moose (Alces alces) habitat and use on glyphosate (nitrogen-phosphonomethyl glycine) (Roundup, Monsanto, St. Louis, Missouri)-treated and untreated clearcuts in north-central Maine. Available browse plants were less abundant on 2-year post-treatment clearcuts compared with untreated clearcuts, except within areas missed during herbicide treatment. Available browse within these skipped areas was similar to that of untreated clearcuts but use was 3.8 times greater in skipped areas. Based on browse use and pellet group distribution, areas of clearcuts missed during treatment were used extensively by Moose. Leaving untreated patches of vegetation in large treated clearcuts is recommended. Schiffman, S. S., M. S. Suggs, M. B. A. Donia, R. P. Erickson, and H. T. Nagle. 1995. Environmental pollutants alter taste responses in the gerbil. Pharmacology, Biochemistry and Behavior 52 , no. 1 : 189-94 . Taste and smell are chemical senses that play a crucial role in food selection. Damage to taste and smell receptors can impair food intake, nutritional status, and survival. The purpose of this study was to determine the effects of 11 environmental pollutants (nine insecticides and two herbicides) on electophysiological taste responses in the gerbil. Integrated chorda tympani (CT) recordings were obtained from gerbils to a range of tastants before and after a 4-min application of 1 of 11 environmental pollutants. The taste stimuli were: sodium chloride (100 mM), calcium chloride (300 mM), magnesium chloride (100 mM), HCl (10 mM), potassium chloride (500 mM), monosodium glutamate (MSG) (50 mM), sucrose (100 mM), fructose (300 mM), sodium saccharin (10 mM), quinine HCl (30 mM), and urea (2 M). The nine insecticides included organophosphorous, carbamate, and pyrethroid insecticides. The seven organophosphorous insecticides tested were: acephate, carbofuran, 89

62.

63.

64.

Mammals chlorpyrifos, chlorpyrifos oxon, demeton, malathion, and methamidophos. The carbamate insecticide carbaryl and the pyrethroid insecticide fenvalerate were also tested. Two herbicides, paraquat and glyphosate, were tested, and dose-response curves for each of these two herbicides were also determined. All of the 11 insecticides and herbicides had an effect on some of the taste stimuli tested. Application of 10 mM methamidophos exhibited the greatest amount of suppression on the 11 taste solutions. Each taste stimulus was significantly suppressed with the exception of 2 M urea. Herbicides paraquat and glyphosate also reduced responses to several tastants. These data indicate that environmental pollutants can modify taste responses in the gerbil. 65. Servello, F. A., B. Griffith, K. S. Raymond, and W. E. Eschholz. 1995. Effects of glyphosate on winter habitat of moose in Maine. College of Natural Resources, Forestry and Agriculture, Maine Agricultural and Forest Experiment Station, University of Maine, Orono, Maine, Cooperative Forestry Research Unit . "CFRU Research Bulletin 10." Misc. Report 395. 19 p. We studied short-term and long-term effects of aerial treatment of clearcuts with the herbicide glyphosate on moose (Alces alces) habitat and activity in winter. For short-term (1-2 years posttreatment) studies, we measured browse availability, use and nutritional quality, winter cover and moose activity on six treated and six untreated clearcuts. For long-term studies (7-11 years posttreatment), we sampled 14 clearcuts that had been treated 7-11 years earlier and five similarly aged untreated clearcuts. In years 1-2 posttreatment, deciduous browse was less abundant (70% in year 2) on treated than untreated clearcuts. There was evidence that red maple (Acer rubrum) and paper birch (Betula papyrifera), two species consistently used by moose on these sites, decreased less than pin cherry (Prunus pensylvanica) and possibly aspen (Populus spp.). Glyphosate had little effect on the nutritional quality of browse available to moose and did not affect the quality of moose diets. The percentage of available browse eaten by moose was not affected by treatment, but counts of tracks of foraging moose were less on treated clearcuts. At 7-11 years posttreatment, treated clearcuts and similarly aged untreated clearcuts had similar amounts of deciduous browse. Treated clearcuts had more browse with a relatively high digestible energy content; however, glyphosate treatment did not affect the nutritional quality of moose diets on these clearcuts. In contrast to effects observed in years 1-2, percentage use of deciduous browse, track counts of foraging moose, and counts of moose beds were greater on treated than untreated clearcuts. We concluded that the reduction in moose activity 1-2 years after treatment was the result of less deciduous browse because we found little evidence for changes in browse or diet quality. Moose appeared to prefer treated clearcuts 7-11 years after treatment because the dense conifer cover allowed foraging and bedding on the same site. We discuss management options for minimizing effects of glyphosate treatment on moose habitat. Smith, E. A., and F. W. Oehme. 1992. The biological activity of glyphosate to plants and animals: A literature review. Veterinary and Human Toxicology 34 , no. 6 : 531-43 . See Plant and Soil Residues Section. Sullivan, T. P. 1990. Demographic responses of small mammal populations to a herbicide application in coastal coniferous forest: population density and resiliency. Canadian Journal of Zoology 68: 874-83. This study was designed to assess the demographic responses of small mammal populations to herbicide-induced habitat alteration in a 7-year-old Douglas-fir plantation near Maple Ridge, British Columbia, Canada. Populations of the deer mouse (Peromyscus maniculatus), Oregon vole (Microtus oregoni), Townsend chipmunk (Eutamias townsendii), and shrews (Sorex spp.) were sampled in control and treatment habitats from April 1981 to September 1983 and from April to October 1985. Recolonization of removal areas by these species was also monitored in both habitats. There was little difference in abundance of deer mice, Oregon voles, and shrews between control and treatment study areas. Chipmunk populations appeared to decline temporarily on the treatment areas relative to controls. Recolonization by voles was not affected by habitat change, but for deer mice was lower on the treatment than control area. Both deer mouse and Oregon vole populations were at comparable densities on control and treatment areas in the second and fourth years after herbicide treatment. The proportion of breeding animals and average duration of life were similar in control and treatment populations of deer mice and voles. These small mammal species should be able to persist in areas of coastal coniferous forest that are treated with herbicide for conifer release.

66.

67.

90

Mammals 68. . 1990. Influence of forest herbicide on deer mouse and Oregon vole population dynamics. Journal of Wildlife Management 54: 566-76. I investigated the direct effects of a forest application of glyphosate herbicide on recruitment, growth, and survival in deer mice (Peromyscus maniculatus) and Oregon voles (Microtus oregoni) by comparing populations in control and treatment habitats. Recruitment of deer mice declined during the first postspray summer and winter, but increased on the treatment area in subsequent years. Generally, there was little difference in recruitment of Oregon voles between control and treatment areas; survival of animals was similar on control and treatment areas. However, female voles did survive significantly (P < 0.05) better on the treatment than on the control area in the postspray summer 1982 and winter 1982-83. Lack of consistent differences in body mass and growth rates of deer mice and voles between control and treatment areas during postspray periods indicated that glyphosate had little or no direct effect on metabolic or general physiological processes in the development of young animals. The manifestation of physiological changes in individual animals that might have resulted from exposure to or ingestion of glyphosate was not apparent in demographic attributes at the population level. . 1994. Influence of herbicide-induced habitat alteration on vegetation and snowshoe hare populations in sub-boreal spruce forest. Journal of Applied Ecology 31 : 717-30 . This study was designed to test the hypothesis that herbicide-induced habitat alteration would reduce snowshoe hare (Lepus americanus) populations in early- to mid-successional (<25 years posttreatment) stages of sub-boreal spruce forest. Intensive population sampling and monitoring of vegetation were conducted in replicate control and treatment blocks near Prince George, British Columbia, Canada from 1988 to 1991. Index line surveys of hares also were conducted in 10 paired control-treatment blocks. Herb biomass and cover recovered to control values by 2-3 years posttreatment. Shrub and tree biomass and cover were little affected in a conifer release treatment where coniferous species dominated these layers. In a backlog conversion treatment, the dominant deciduous trees and shrubs were relatively slow to recover. Herbicide-induced habitat alteration in optimum habitat seemed not to affect abundance of snowshoe hares during summer and autumn. Post-harvest forest habitats 10-20 years of age may have sufficient and persistent vegetative structure, particularly those with a high component of coniferous species, to support hare populations regardless of herbicide treatment. . 1996. Influence of forest herbicide on snowshoe hare population dynamics: reproduction, growth, and survival. Canadian Journal of Forest Research 26 , no. 1 : 112-19 . This study was designed to assess the influence of forest applications of glyphosate herbicide on reproduction, growth, and survival in snowshoe hare (Lepus americanus Erxleben) populations in control (reference) and treatment habitats near Prince George, B.C. Proportion of adult hares in breeding condition and number of successful pregnancies showed no consistent differences between control and treatment populations. Recruitment of hares was generally similar except for significantly more juvenile females entering the control than treatment population at one study area. At a second study area, total recruitment was significantly higher in the treatment than control population for both sexes in 1990 and for adult females in 1991, the 2 post-treatment years. There was little difference in survival of hares between control and treatment populations. Lack of significant differences in mean body mass and growth rates suggested that this herbicide treatment had little or no effect on metabolic or general physiological processes in the development of young hares. Similar profiles of body mass distribution between control and treatment populations indicated that comparable levels of biomass of hares were available as prey for predators. Use of this forest herbicide did not measurably affect demographic parameters of snowshoe hare populations. Sullivan, T. P., and J. O. Boateng. 1996. Comparison of small-mammal community responses to broadcast burning and herbicide application in cutover forest habitats. Canadian Journal of Forest Research 26 : 462-73 . This study was designed to compare the responses of small-mammal communities to broadcast burning and herbicide-induced alteration of forest habitats. Study areas were located in south-coastal British Columbia, Canada, in the Coastal Western Hemlock (CWHdm) biogeoclimatic zone, and in west-central British Columbia in the Sub-boreal Spruce (SBSmk) and Engelmann Spruce91

69.

70.

71.

Mammals Subalpine Fir (ESSFmc) zones. Control-treatment comparisons for a herbicide application or a broadcast burning treatment were conducted at the coastal study area from 1982 to 1984. Replicate control-treatment comparisons between these two silvicultural practices were conducted at the interior study area within the period 1989 to 1992. Small-mammal populations were intensively livetrapped in all control and treatment blocks. Deer mouse (Peromyscus maniculatus Wagner) populations showed short-term (1-2 months) declines after treatments at the coastal study area but appeared little affected by these habitat alterations at the interior area. Voles of the genus Microtus disappeared from burned blocks at the interior area but Microtus oregoni (Merriam) persisted on the burned block at the coastal area; however, the red-backed vole (Clethrionomys gapperi Vigors) did not. Chipmunks (Eutamias townsendii Bachman and Eutamias amoenus Allen) were little affected by either treatment. Neither treatment seemed to affect species diversity of these small-mammal communities. In terms of abundance of small-mammal populations, it is likely that broadcast burning is a more extreme means of habitat alteration than herbicide treatment. 72. Sullivan, T. P., and E. J. Hogue. 1987. Influence of orchard floor management on vole and pocket gopher populations and damage in apple orchards. Journal of the American Society for Horticultural Science 112: 972-77. Removal of all vegetation with herbicides over the total orchard floor or only in tree rows significantly reduced montane vole (Microtus montanus Peale), meadow vole (M. pennsylvanicus Ord), and northern pocket gopher ( Thomomys talpoides Richardson) populations and damage. Herbicide treatments in four test orchards were carried out during May, July, and Sept. 1983 to 1985. Average overwinter abundance of voles was reduced 53% to 99% on treatment areas. Several vole populations went to extinction in the third year of herbicide treatment. Incidence of tree damage was 40.6% and 9.6% with feeding intensities of 17.2 cm2 and 0.4 cm2 of bark and tissues removed per tree on control and treatment blocks, respectively, during a peak year in abundance of voles. Pocket gopher populations and damage were significantly lower in treatment than control blocks. Deer mouse (Peromyscus maniculatus Wagner) and yellow pine chipmunk (Eutamias amoenus J.A. Allen) populations generally increased on treated areas. Use of herbicides to control orchard floor vegetation is an effective means of rodent damage control. Sullivan, T. P., R. A. Lautenschlager, and R. G. Wagner. 1995. Changes in diversity of plant and small mammal communities after herbicide application in sub-boreal spruce forest. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R.E. Gaskin, J.A. Zabkiewicz, (compilers). FRI Bulletin No. 192. pp. 143-45. This study evaluated changes in species diversity of plant and small mammal communities to herbicide application for conifer release in sub-boreal spruce forest. There were no differences in number of plant species between control and treatment blocks in the herb (5-year-old) successional stage, but shrub species richness declined in the shrub (10-year-old) successional stage. There were no differences in diversity of herbs throughout the study. Diversity of shrubs was lower in posttreatment than control blocks in both successional stages, but primarily in the first two years after treatment. Species diversity of small mammal communities was not affected. Except for a temporary decline in shrub richness and diversity, these treatment blocks should not lower overall diversity of a forested landscape. Sullivan, T. P., C. Nowotny, R. A. Lautenschlager, and R. G. Wagner. 1997. Silvicultural use of herbicide in sub-boreal spruce forest: implications for small mammal population dynamics. (Submitted). This study was designed to test the hypothesis that herbicide-induced habitat alteration would reduce small mammal populations in early successional stages of sub-boreal spruce forest. A secondary objective was to determine the long-term influence of herbicide treatment on reproduction, survival, and growth attributes of deer mouse (Peromyscus maniculatus) and southern red-backed vole (Clethrionomys gapperi) populations. Small mammal populations were intensively sampled in 4 replicate paired control and treatment sites of herb-shrub successional stages (5-10 years after harvesting) near Prince George, British Columbia, Canada. Sampling was conducted during 1987 (pretreatment year), 1988-1989 (first and second posttreatment years), and in 1991-1992 (fourth and fifth posttreatment years). Shrub layers were significantly reduced on all study areas after treatment. 92

73.

74.

Mammals Herb layers recovered to pretreatment or control levels of biomass in the second posttreatment year. Average numbers of red-backed voles and shrews (Sorex spp.) were consistently higher on control than treatment sites during 1988 to 1992. Average abundance of meadow voles (Microtus pennsylvanicus) and deer mice were similar on control and treatment sites throughout the study. Weasels (Mustela erminea and Mustela frenata) appeared on those control and treatment sites which supported prey populations of mice and voles. There were no consistent differences in proportion of adult male and female red-backed voles and deer mice in breeding condition between control and treatment populations during the study. However, there was a significantly higher average number of successful pregnancies in control than treatment populations of red-backed voles during posttreatment years. Deer mouse populations showed no difference in this measure of reproductive performance between control and treatment sites. Average estimates of Jolly survival of red-backed voles were significantly higher in treatment than control populations, with no difference in average survival of deer mice between control and treatment populations. Herbicide treatment had no effect on mean body mass of male or female red-backed voles or deer mice. Total biomass of rodents was similar in the small mammal communities occupying control and treatment sites. This is the first relatively long-term (up to 5 years posttreatment) investigation of small mammal responses to herbicide application in replicated experimental units in northern coniferous forest, a major ecological zone in North America and Eurasia. Herbicide treatment of areas dominated by early successional vegetation in northern coniferous forests should be staggered in time and space to allow recovery (2-3 years) of vegetation and small mammal species such as red-backed voles and shrews. A mosaic of different-aged forests, with respect to harvesting history and herbicide treatment, would be optimum for maintaining a variety of habitats and small mammal species. 75. Sullivan, T. P., and D. S. Sullivan. 1979. The effects of glyphosate herbicide on food preference and consumption in black-tailed deer. Canadian Journal of Zoology 57: 1406-12. The use of herbicides is an important part of forestry management practices in the Pacific Northwest because the regeneration of coastal forests is hampered by many species of deciduous shrubs and weeds. The herbicide glyphosate is used to control these undesirable species. Some effects of glyphosate on black-tailed deer have been investigated by analyzing food preference and consumption under simulated field conditions. Deer given a choice of control or glyphosate-treated alder and alfalfa browse showed no preference or ate more of the treated foliage. The ingestion of treated browse did not affect the consumption of laboratory chow by the deer. These results indicate that spraying with the herbicide glyphosate should not prevent deer from feeding on foliage in the affected area. . 1981. Responses of a deer mouse population to a forest herbicide application: reproduction, growth, and survival. Canadian Journal of Zoology 59: 1148-54. This study was designed to monitor some of the demographic responses of a deer mouse population to a forest application of RoundupR herbicide. Populations of Peromyscus maniculatus were livetrapped from July 1978 to November 1980 on a control area and in a herbicide-treated 20-year-old Douglas-fir plantation at Maple Ridge, British Columbia. The herbicide had no apparent adverse effects on reproduction, growth, or survival of deer mice 1 year after treatment. Inconsistencies in growth rates and juvenile survival between control and experimental deer mice in 1980 could be due to the herbicide or demographic factors. Field dose applications of this herbicide should not have a direct effect on the dynamics of deer mouse populations. . 1982. Responses of small-mammal populations to a forest herbicide application in a 20year-old conifer plantation. Journal of Applied Ecology 19: 95-106. The responses of small-mammal populations in a forest application of RoundupR herbicide have been investigated at the University of British Columbia Research Forest, Maple Ridge, B.C., Canada. These populations included the deer mouse, Oregon vole, Townsend chipmunk, and shrews. Treatment of a 20-year-old Douglas fir plantation did not have any negative effects on the distribution and abundance of small-mammal populations during the first year after this habitat alteration. Movements of deer mice were monitored by drift lines. There was not an influx of new animals from the surrounding regions onto the treated area nor was there a significant movement of marked animals

76.

77.

93

Mammals away from the sprayed area. Future changes in composition of the small-mammal community may occur in association with successional stages advancing from the herbicide-induced habitat alteration. 78. Sullivan, T. P., D. S. Sullivan, E. J. Hogue, R. A. Lautenschlager, and R. G. Wagner. 1997. Population dynamics of small mammals in relation to habitat alteration in orchard agroecosystems: Compensatory responses in abundance and biomass. (Submitted). See Biodiversity and Habitat Restoration Section. Sullivan, T. P., D. S. Sullivan, R. A. Lautenschlager, and R. G. Wagner. 1997. Long-term influence of glyphosate herbicide on demography and diversity of small mammal communities in coastal coniferous forest. Northwest Science 71 , no. 1 : 6-17 . This study tested the hypothesis that (1) abundance and related demographic parameters of small mammal populations, and (2) species diversity of small mammal communities, would be adversely affected in herbicide-treated habitats at 9 and 11 years after treatment in coastal coniferous forest. Study areas were located in south-coastal British Columbia, Canada, in the Coastal Western Hemlock (CWHdm) biogeoclimatic zone where small mammal populations were intensively sampled on paired control and treatment sites. Average density of deer mice (Peromyscus maniculatus) during summer periods (1981-1990) was lower in the treatment than control in the immediate post-treatment (PT) period (1982), with comparable numbers in 1983, 1985, and 1990. At the 11-year PT area, deer mouse numbers were similar on control and treatment sites in 1978, 1979, 1980, and 1990. Average density of Oregon voles (Microtus oregoni) was higher on treatment than control sites at 9 and 11 years after treatment. Townsend chipmunks (Eutamias townsendii) tended to be less abundant on the treatment than control site 9 years PT but were essentially absent from the 11-year PT study area. There was no difference in average density of shrew (Sorex spp.) populations between control and treatment sites at either study area. It is likely that post-harvest successional change has more of an impact on small mammal abundance than change induced by a herbicide treatment. Our results suggest that glyphosate herbicide did not adversely affect reproduction, survival, or growth of deer mice and Oregon voles in coastal forest a decade after application. Species richness and diversity of small mammal communities changed little over the decade following treatment. This study is the first investigation on the effects of forest herbicide use on demography and diversity of small mammal communities that extends to a decade. Sullivan, T. P., and L. W. Taylor. 1995. Wildlife habitat enhancement with Ezject in stand thinning of lodgepole pine. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R.E. Gaskin, J.A. Zabkiewicz (compilers). FRI Bulletin No. 192. pp. 310-312 . See Biodiversity and Habitat Restoration Section. Sullivan, T. P., R. G. Wagner, D. G. Pitt, R. A. Lautenschlager, and D. G. Chen. 1997. Changes in diversity of plant and small mammal communities after herbicide application in sub-boreal spruce forest. (Submitted) . See Biodiversity and Habitat Restoration Section. Thompson, D. G., D. G. Pitt, B. Staznik, N. J. Payne, D. Jaipersaid, R. A. Lautenschlager, and F. W. Bell. 1997. On-target deposit and vertical distribution of aerially released herbicides. Forestry Chronicle 73: (In press) . See Plant and Soil Residues Section. Timmerman, H. R., J. G. McNicol, and C. S. Krishka. 1986. Impact of glyphosate on wildlife habitat: A three year study. Ministry of Natural Resources, Wildlife and Forest Resources Branch, Government of Ontario, Ontario. This study was designed to establish the impact of aerial application of glyphosate for conifer release (in the boreal forest of northern Ontario) by studying the changes in vegetative composition and density, and its effect on the winter utilization by moose and snowshoe hare and fall utilization by small mammals, over a three year study period. Treated and adjacent untreated areas are being studied and compared to determine differences, if any, due to herbicide treatment. The study methodology is 94

79.

80.

81.

82.

83.

Mammals designed to assess changes over a three year period in the vegetative community on areas treated with glyphosate through comparison with similar untreated areas. Changes in the vegetative community may affect the utilization of treated sites by herbivores (moose and small mammals) and thus carnivores (terrestrial furbearers). Periodic assessment would be beneficial after the initial threeyear study to observe long term changes, i.e. 5, 10 and 15 years. 84. Trichet, P., B. Boisaubert, H. Frochot, and J. F. Picard. 1987. Impact of herbicide treatments against bramble Rubus fruticosus Lagg on roe deer Capreolus capreolus L. Gibier Faune Sauvage 4: 165-88. The effects of a herbicide treatment on the fauna are two-fold: the direct, or indirect, toxic effects which are studied before a product obtains approval for registration and the herbicidal sideeffects due to phenomena of substitution among plants. As a matter of fact, modifications in the food availability and the habitat structure do affect the behaviour of animal populations. Therefore, the impact of treatments against bramble (Rubus fruticosus Lagg) with glyphosate herbicide was studied in Lorraine (France). Bramble being one of the staple foods of Roe deer in winter, the use of treated and untreated plots by Roe deer was compared. From the forester's point of view, the effectiveness of the treatments is generally satisfactory, although the brambles are not entirely destroyed. At Lorraine the Roe deer densities of readily 6-10 animals per ha, the treatment does not change the intensity in Roe deer use. Browsing pressure is somewhat higher on treated plots. In case where brambles are 80% destroyed and population densities are high (30 to 35 animals per 100 ha) Roe deer change their feeding behaviour by foraging on shrubs instead of brambles. This phenomenon does not persist: 4 to 5 years after treatment, bramble bushes have recovered and Roe deer do not seem to distinguish between control and herbicide-treated plots any more. (France). Wahlgren, J. R. 1979. "The effects of the herbicide Roundup on reproduction in mice. " B.S.Agr. Thesis, Dept. of Animal Science, University of British Columbia. The herbicide RoundupR has the potential for increased use in the control of competitive shrub and tree species in forest plantations. This report examines the effect of RoundupR on reproductive performance in mammals which may inhabit these forest ecotypes. Laboratory mice are used as test animals. For the variable chosen to represent the physiological processes of reproduction, no differences were found using F-tests and Chi-squared analyses at the five per cent probability level. The experiment was run over a period of two litters from the same parents. William, S. I. 1995. Ten-year development of Douglas-fir and associated vegetation after different site preparation on Coast Range clearcuts. U.S.D.A. Forest Service Research Paper PNW 0 473 . Yousef, M. I., M. H. Salem, H. Z. Ibrahim, S. Helmi, M. A. Seehy, and K. Bertheussen. 1995. Toxic effects of carbofuran and glyphosate on semen characteristics in rabbits. Journal of Environmental Science and Health-B 30 , no. 4 : 513-34. See Human Health Section.

85.

86.

87.

95

Microflora and Fungi

Microflora and Fungi


1. Abdel-Mallek, A. Y., M. I. A. Abdel-Kader, and A. M. A. Shonkeir. 1994. Effect of glyphosate on fungal population, respiration and the decay of some organic matters in Egyptian soil. Microbiological Research 149, no. 1 : 69-73. Glyphosate (Roundup), when applied to the soil usually did not exert any significant effect on the total count of soil fungi after all periods of the experiment except after 6 and 10 weeks where the count was inhibited by the two doses used (1.84, 9.2 mg active ingredient/kg dry soil). When the herbicide was incorporated into the agar medium the count of total fungi, Acremonium strictum and Aspergillus fumigatus was significantly increased by the two doses used and of Penicillium glabrum by the high dose only. However, P. funiculosum was completely eliminated by the high dose. Oxygen consumption in soil treated with glyphosate was significantly inhibited by the high dose after 2 weeks and by the two doses after 6, 8 and 10 weeks. Glyphosate exerted two significant effects of stimulation and inhibition on the rate of the decay of stem segments of three plants at certain treatments of dose and time. (Egypt). Ana'yeva, N. D., B. P. Strekozov, and O. K. Tyuryukanova. 1987. Change in the microbial biomass of soils caused by pesticides. Soviet Soil Science 18: 56-62. (Soviet Union). Aubin, A. J., and A. E. Smith. 1992. Extraction of Carbon-14-labelled glyphosate from Saskatchewan soils. Journal of Agriculture and Food Chemistry 40: 1163-65. The extraction of 14C from air-dried samples of three soils fortified 14 days previously with (14C)-glyphosate was compared using 10 different solvent systems and an extended-shaking procedure. 14C recoveries exceeding 73% were reproducibly achieved with 0.35 M H3PO4 (and 0.09 M with respect to CaCl2), 0.1 M NaOH, and 0.5 M NH4OH. With the phosphoric acid and ammonium hydroxide extractants highest recoveries from a loamy sand were recorded with a soil/solvent ratio of 1:5 or 1:10. With the aqueous sodium hydroxide, highest recoveries were achieved from all soils using a soil/solvent ratio of 1:2.5. Recoveries of 14C from all three soils, using 0.1 M NaOH, were not significantly different whether fortified 14, 28, or, in the case of clay, 56 days previously with (14C)glyphosate. Some loss of extractable radioactivity was noted in two of the soils after 56 days. Bergvinson, D. J., and J. H. Borden. 1992. Enhanced colonization by the blue stain fungus Ophiostoma clavigerum in glyphosate-treated sapwood of lodgepole pine. Canadian Journal of Forest Research 22: 206-9. The herbicide glyphosate was administered into the sapwood around the root collar of lodgepole pine trees, Pinus contorta var. latifolia Engelm., to determine its effect on invasion by the blue stain fungus Ophiostoma clavigerum (Robinson-Jeffrey & R.W. Davidson) T.C. Harrington. In two experiments, lesions in the sapwood were longer and wider in trees treated with glyphosate before inoculation with O. clavigerum than in untreated, control trees. Ophiostoma clavigerum was recovered in a third experiment at seven times the distance from the point of inoculation in trees treated with glyphosate 3 weeks before inoculation as in untreated, control trees. We conclude that previously observed enhancement of brood development of the mountain pine beetle, Dendroctonus ponderosae Hopk., was caused by glyphosate-induced inhibition of the trees' secondary defense response to invasion by the beetle's symbiotic fungi. Berner, D. K., G. T. Berggren, and J. P. Snow. 1991. Effects of glyphosate on Calonectria crotalariae and red crown rot of soyabean. Plant Disease 75: 809-13. Surfactant- and nonsurfactant-containing commercial herbicide formulations of glyphosate were evaluated for in vitro and in vivo effects on C. crotalariae and red crown rot of soyabean, respectively. Rates of 0.28, 0.56, 1.12 and 2.25 kg of glyphosate per hectare, corresponding to recommended rates for weed control in soyabean were used. Three pathogenic isolates of the fungus from soyabean were grown on a selective medium amended with either water or the different rates of the herbicide. Both formulations of glyphosate inhibited mycelial growth by C. crotalariae. Additions of

2.

3.

4.

5.

96

Microflora and Fungi amino acids to medium amended with the nonsurfactant formulation produced a reversal of herbicide inhibition. Duplicated field trials showed a reduction in red crown rot incidence with preplant applications of low rates of glyphosate. The results, coupled with findings on the significant influence of previous season disease incidence on current red crown rot levels, indicate that glyphosate may be used simultaneously and efficaciously as a preplant herbicide for weed control and as a fungicide for the control of diseases caused by C. crotalariae. 6. Beyrle, H. F., S. E. Smith, R. L. Peterson, and C. M. M. Franco. 1995. Colonization of Orchis morio protocorms by a mycorrhizal fungus: Effects of nitrogen nutrition and glyphosate in modifying the responses. Canadian Journal of Botany 73, no. 8 : 1129-40. Effects of nitrogen nutrition and application of glyphosate (as Roundup) on the interactions between protocorms of Orchis mori and a mycorrhizal Rhizoctonia species were investigated. Protocorms for all experiments were raised clonally in liquid culture. Split plates were used to separate the direct effects of composition of the media from effects mediated via the fungus. Mycorrhizal interactions, including coil formation and prolonged protocorm growth, were established when a relatively low nitrogen supply to the fungus (as NH-4)-2SO-4) was combined with a high carbohydrate supply. Rejection of the fungus, associated with phenolic production, wall thickening, and lack of protocorm growth, was observed with high-carbon, high-nitrogen medium. Low carbohydrate supply, with either high or low nitrogen supply, was associated with breakaway parasitism, symptoms of soft rot, and lack of subsequent protocorm growth. Application of glyphosate at 0.5 or 1.0 mM had no effect on fungal growth and at 1.0 mM did not cause death of asymbiotic protocorms, but resulted in failure of mycorrhizal initiation. Coils were never formed and the fungus ramified through the tissues of the protocorms. The parasitism induced by glyphosate differed from breakaway parasitism, and there were no symptoms of soft rot. Assays for activity of phenylalanine ammonia lyase (PAL) and for orchinol were carried out to assess the effects of the treatments on the shikimic acid pathway. Low activity of PAL and low quantities of orchinol (together with another unidentified phenolic compound) were detected in asymbiotic protocorms. Concentrations of both were increased in the presence of the fungus, but no significant differences were observed in the various symbiotic responses in the absence of glyphosate. This is the first report of the presence of orchinol in orchid protocorms. Previous work has always been carried out on tubers, which are not usually colonized by the mycorrhizal fungi. Application of glyphosate resulted in increases in both PAL activity and orchinol production. The results are discussed in the context of orchid-fungus interactions and the mechanism of glyphosate action. (Australia) . Bezbaruah, B., N. Saikia, and T. Bora. 1995. Effect of pesticides on most probable number of soil microbes from tea (Camella sinensis) plantations and uncultivated land enumerated in enrichment media. Indian Journal of Agricultural Sciences 65 , no. 8 : 578-83 . To enumerate the most probable number (MPN) of soil microbes from tea (Camellia sinensis (L.) O. Kuntze) plantations and uncultivated land. Soil samples were placed into enrichment media designed to promote cellulolytic, nitrogen-fixing or sulphur-oxidizing groups. The experiment was conducted in 1990-91 by collecting fresh soil samples 6 times a year to cover all the seasonal variations. To study the effect of pesticides, each sample was introduced into soil dilutions 24 hr before plating into the respective enrichment medium. The population of sulphur-oxidizing strains increased in the presence of malathion and dimethoate and reduced in the presence of fenitrothion and endosulfan. All the pesticides tried reduced the population of nitrogen-fixing strains. Glyphosate and 2,4-D increased the population of cellulolytic strains but dalapon and paraquat reduced it. (India) . Bliev, Y. K., and A. N. Martynov. 1981. The influence of glyphosate on the fertility of a sodpodsolic soil. Khimiya v Sel'Skom Khozyaistive 19: 51-54. (Soviet Union). Bode, R., F. Schauer, and D. Birnbaum. 1986. Comparative studies on the enzymological basis for growth inhibition by glyphosate in some yeast species. Biochemie Und Physiologie Der Pflanzen 181: 39-46. We examined the mode of the action of the herbicide glyphosate (N-(phosphonomethyl)glycine) on 19 strains of Ascomycetous and Basidiomycetous yeasts (Brettanomyces, Candida,

7.

8.

9.

97

Microflora and Fungi Cryptococcus, Hansenula, Kluyveromyces, Lodderomyces, Pichia, Rhodosporidium, Rhodotorula, Saccharomycopsis, Schizosaccharomyces, Sporobolomyces, Yarrowia). In all yeast species we observed the following physiological and enzymological effects: (1) aromatic amino acids reverse the glyphosate-induced growth inhibition, (2) yeast cells excrete shikimic acid in the presence of glyphosate, (3) the tyrosine-sensitive enzyme was the only one of the 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase isoenzymes which was inhibited by glyphosate at concentrations in the millimolar range, (4) 5-enolpyruvylshikimate 3-phosphate (EPSP) synthase activity was inhibited by micromolar concentrations of glyphosate. Yarrowia lipolytica was extremely sensitive, whereas a similar growth inhibition of Candida albicans, C. utilis and Hansenula polymorpha was only achieved by about 400fold concentration of the herbicide. Our in vivo results show that the presumable primary site of the glyphosate action in all yeast strains is the effective inhibition of the EPSP synthase activity. (Germany). 10. Bogdanovic, V. V. 1975. Study of the herbicide effects upon some soil microorganisms. Mikrobiologija 12: 121-25. Spraying the soil of an apple orchard with 10 kg simazine/ha did not affect the soil microorganisms. Caragard increased the total numbers of microorganisms, and especially those of fungi and actinomycetes. Roundup increased the total microorganisms and aerobic ammonifiers, but inhibited fungi and actinomycetes. (Yugoslavia). Bora, T. K., and B. Bezbaruah. 1986. Characterization and pesticide concentration tolerance of cellulolytic bacteria from tea soils. Industrial Journal of Microbiology 26: 278-87. (India). Bujacz, B., P. Wieczorek, T. Krzysko-Lupicka, Z. Golab, B. Lejczak, and P. Kavfarski. Organophosphonate utilization by the wild-type strain of Penicillium notatum. Applied and Environmental Microbiology 61 , no. 8 : 2905-10. We studied the biodegradation of compounds containing phosphorus-to-carbon bonds by using a wild-type strain of Penicillium notatum. The substrate specificity of this strain was studied, and we found that it is able to utilize structurally diverse organophosphonates as sole sources of phosphorus. This ability seems to be inducible, as indicated by the presence of a lag phase during growth. A popular herbicide, glyphosate, inhibited fungal growth, but it was also degraded by the fungus if it was applied in sublethal doses. This indicates that P. notatum may play an important role in biodegradation of organophosphonates. The strain which we used did not metabolize any of the phosphonates which we tested when they were used as sole carbon or nitrogen sources. (Poland) . Carlisle, S. M., and J. T. Trevors. 1986. Effect of the herbicide glyphosate on nitrification denitrification and acetylene reduction in soil. Water, Air, and Soil Pollution 29: 189-204. The effect of glyphosate on soil respiration and H2 oxidation in an agricultural soil was R investigated. The effects of the pure herbicide and commercial formulation, Roundup (Monsanto Company), were compared in soil under both aerobic and anaerobic conditions. Both formulations stimulated 02 uptake as well as aerobic and anaerobic C02 evolution. Roundup caused more stimulation than glyphosate under aerobic incubation conditions; the formulations had an equal effect on anaerobic C02 evolution. Hydrogen oxidation was inhibited by both formulations in aerobic and anaerobic soil. Aerobic oxidation was inhibited to the same extent by both formulations; Roundup had a stronger inhibitory effect on anaerobic H2 oxidation than did glyphosate. No toxicity to any of these activities should be seen at recommended field application rates of the herbicide. . 1988. Glyphosate in the environment. Water, Air, and Soil Pollution 39 : 409-20. See Plant and Soil Residues Section. Chakravarty, P., and L. Chatarpaul. 1990. Non-target effect of herbicides. I. Effect of glyphosate and hexazinone on soil microbial activity, microbial population and in vitro growth of Ectomycorrhizal fungi. Pesticide Science 28: 233-42. The effects of two herbicides, glyphosate (as a 359 g/l SL) and hexazinone (as a 50 g/kg granules) on soil microbial population, carbon dioxide evolution, and in vitro growth of five species of 98

11.

12.

13.

14.

15.

Microflora and Fungi ectomycorrhizal fungi were investigated. Glyphosate at 0.54 and 3.23 kg a.i./ha and hexazinone at 1, 2 and 8 kg a.i./ha did not reduce soil microbial population or carbon dioxide evolution in the long term (6 months). However, there was a significant short-term (2 months) effects of glyphosate on both fungal and bacterial counts at the 0.54 kg/ha treatment. In in vitro tests, Cenococcum graniforme, Hebeloma crustuliniforme and Laccaria laccata were more susceptible to both herbicides than was Suillus tomentosus, which was, in turn, more susceptible than Paxillus involutus. The growth of all five ectomycorrhizal fungi was significantly reduced when subjected to concentrations above 50 m.l formulation/litre (glyphosate) or 50 m.g formulation/litre (hexazinone). 16. . 1990. Non-target effect of herbicides. II. The influence of glyphosate on ectomycorrhizal symbiosis of red pine Pinus resinosa under greenhouse and field conditions. Pesticide Science 28: 243-48. The effects of the herbicide glyphosate (N-(phosphonomethyl) glycine) on seedling growth of Pinus resinosa and on the ectomycorrhizal development of the symbiotic fungus Paxillus involutus were investigated under greenhouse and field conditions. Glyphosate at 0.54 and 3.23 kg/ha did not reduce seedling growth or ectomycorrhizal development under greenhouse conditions. There was no seedling mortality due to glyphosate treatment at either rate. Under field conditions, seedling growth of P. resinosa and ectomycorrhizal development of P. involutus were not affected by the above rates. Seedlings (46-48%) failed to survive in non-glyphosate-treated plots, presumably because of weed competition, whereas all survived in glyphosate-treated plots. All the non-mycorrhizal seedlings were colonized by indigenous mycorrhizal fungi within 2 months after planting in both control and glyphosate-treated plots. The infection rates varied from 74-86%. Chakravarty, P., and S. S. Sidhu. 1987. Effect of glyphosate, hexazinone and triclopyr on in vitro growth of five species of ectomycorrhizal fungi. European Journal of Forest Pathology 17: 204-10. In vitro growth tests with glyphosate (Roundup), hexazinone (liquid Velpar L.R and granule PrononetM5GR) and trichlopyr (GarlonR) on five species of ectomycorrhizal fungi (Hebeloma crustuliniforme, Laccaria laccata, Thelophora americana, T. terrestris and Suillus tomentosus) showed varied species sensitivity to different concentrations of herbicides. Fungal growth was significantly (P = 0.05) reduced particularly at concentrations above 10 ppm. Garlon with trichlopyr as a.i. (active ingredient) was the most toxic of the four herbicide formulations. Chan, K., and S. C. Leung. 1986. Effects of paraquat and glyphosate on growth, respiration, and enzyme activity of aquatic bacteria. Bulletin of Environmental Contamination and Toxicology 36: 52-59. (*) The present paper reports the results of a study on the effects of paraquat and glyphosate on growth, respiration and enzyme activity of aquatic bacteria. Experimental pools A, B and C were located at the Lam Tseun River in Hong Kong. All these pools were gravel pits with surface areas of 25, 23 and 21 m2, and mean depths of 0.8, 0.6 and 0.54 m, respectively. Paraquat (GramoxoneR, 1,1'dimethyl-4,4'-bipyridinium dichloride) was applied to pool B so a concentration of 20 ppm was obtained. Glyphosate (RoundupR, isopropylamine salt of N-phosphonomethyl glycine containing 356 g/l glyphosate) was applied to pool C so that a concentration of 200 ppm was achieved. Pool A was used as the control. Viable cell counts of the pool water were determined at 2-day intervals by using the spread plate method with Nutrient Agar for a total of 30 days after application of the herbicides. All agar o plates were incubated at 25 C for 48 hours before the number of bacterial colony forming unit (CFU) was counted. Two bacteria of most frequent occurrence were identified as Aeromonas hydrophila C-1 which was a facultative anaerobic strain, and Pseudomonas chlororaphis L-1 which was a strictly aerobic strain. These two bacteria were used in subsequent experiments. The present results indicate that aquatic bacteria in general are more susceptible to paraquat than the soil microflora and human bacteria in terms of paraquat concentration applied to the aquatic environment. Since in situ study confirms the inhibitory effects of these two herbicides on aquatic bacterial populations and it is conceivable that communities of higher trophic level in the aquatic environment might also be affected. Consequently, consideration must be given to the possibility of contamination of non-target environments such as streams, ponds and lakes when applying paraquat and glyphosate in the field. (Hong Kong). 99

17.

18.

Microflora and Fungi 19. Dalla-Chiesa, M., S. R. Mayes, D. J. Maskella, P. J. Nixon, and J. Barber. 1994. An aroA homologue from Synechocystis sp. PCC6803. Gene (Amsterdam) 144, no. 1 : 145-46. In this study, we report the entire nucleotide sequence of an aroA homologue encoding 5enolpyruvylshikimate-3-phosphate synthase (EPSPS), isolated from the cyanobacterium Synechocystis sp. PCC 6803. The proposed coding region is an open reading frame of 447 amino acids. The deduced sequence of the gene product is particularly similar to the Gram+ EPSPS sequences available to date, in particular to that in Bacillus subtilis. Analysis of the Synechocystis putative EPSPS sequence does not lead to an obvious explanation for the natural tolerance of this cyanobacterium to glyphosate. (United Kingdom) . Defelice, M. S., and J. C. Henning. 1990. Renovation of endophyte Acremonium coenophialum infected tall Fescue Festuca-arundinacea pastures with herbicides. Weed Science 38: 629-33. Fall, spring, or late-summer applications of glyphosate, paraquat, sethoxydim, and HOE-39866 were investigated in field trials for no-tillage renovation of endophyte fungus-infected tall fescue pastures. Only spring or late-summer applications of glyphosate at 1.68 and 2.52 kg a.e./ha consistently provided greater than 80% visual tall fescue control. However, none of the herbicide treatments significantly reduced final percent endophyte fungus infection levels after the pasture was reestablished. High final infection levels were due to regrowth from underground rhizomes of old tall fescue sod that survived the herbicide treatments. Descalzo, R. C., C. A. Levesque, Z. K. Punja, and J. E. Rahe. 1992. Influence of fast and slowgrowing Pythium species on the herbicidal effect of glyphosate on beans. Phytopathology 82: 11-15. Pythium isolates from roots of glyphosate-treated bean plants grown in different soil types were identified and grouped based on DNA RFLP patterns. Each isolate was tested for growth rate on PDA and pathogenic characteristics on beans. Colony growth rate was evaluated after 24 hours, and pathogenicity was tested by planting bean seeds in infested metro-mix soil medium to determine the rate of pre-emergence damping-off. Representative isolates from each group were compared for their efficiency as glyphosate synergists. This was done by treating 2-week-old bean plants growing in soil medium infested with the respective isolate with six different doses of glyphosate (0, 2, 6, 20, 50, and 150 g/plant). Two weeks after treatment, plant mortality was counted and the LD50 was determined by logistic regression. Pathogenicity tests showed that fast-growing isolates were more virulent than the slow-growing isolates, and lower glyphosate levels were needed to kill plants grown in soil infested with the former than with the latter. These results suggest that growth rates of Pythium may reflect their efficiency as glyphosate synergists. Dick, R. E., and J. P. Quinn. 1995. Control of glyphosate uptake and metabolism in Pseudomonas sp. 4ASW. FEMS Microbiology Letters 134, no. 2-3 : 177-82 . An environmental bacterial isolate, Pseudomonas sp. 4ASW, was shown to metabolize the organophosphonate herbicide glyphosate as sole phosphorus source, through an initial cleavage of the carbon-phosphorus bond to yield sarcosine. Glyphosate metabolism was repressed in the presence of inorganic phosphate; levels of uptake of the herbicide in cells subjected to prolonged phosphorus starvation were up to 20-fold higher than those in cells actively growing on glyphosate. In contrast, alkaline phosphatase activity was only two-fold higher in P-starved cells. This suggests that there may be significant differences in the regulation of these two phosphate-starvation-inducible functions. (United Kingdom) . . 1995. Glyphosate-degrading isolates from environmental samples: Occurrence and pathways of degradation. Applied Microbiology and Biotechnology 43, no. 3 : 545-50 . The metabolism of the organophosphonate herbicide glyphosate was investigated in 163 environmental bacterial strains, obtained by a variety of isolation strategies from sites with or without prior exposure to the compound. Isolates able to use glyphosate as sole phosphorus source were more common at a treated site, but much less abundant than those capable of using the glyphosate metabolite aminomethylphosphonic acid (AMPA). Nevertheless, all 26 strains found to metabolise the herbicide did so via an initial cleavage of its carbon-phosphorus bond to yield sarcosine; no evidence for its metabolism or co-metabolism to AMPA was obtained. (United Kingdom) . 100

20.

21.

22.

23.

Microflora and Fungi 24. Estok, D., B. Freedman, and D. Boyle. 1989. Effects of the herbicides 2,4-D, glyphosate, hexazinone and triclopyr on the growth of three species of ectomycorrhizal fungi. Bulletin of Environmental Contamination and Toxicology 42: 835-39. (*) This paper reports the results of studies of the toxicity of four herbicides to representative species of ectomycorrhizal fungi that infect forest trees. The herbicides 2,4-D and glyphosate are presently used in silviculture in Canada, while triclopyr and hexazinone are unregistered for forestry purposes. All four of the herbicides are registered for forestry purposes in the United States. Each herbicide significantly reduced the radial growth of each species of ectomycorrhizal fungus at concentrations >1000 ppm. Growth was completely inhibited at concentrations >5000 ppm. In our bioassays, we found evidence of inhibition of fungal growth at herbicide concentrations <100 ppm. However, it is important to note that agar medium presents a very different bioassay condition from that experienced at a similar bulk concentration of herbicide in the much more complex and variable environment of the forest floor or soil. In general however, the conditions in agar tend to predispose the bioassay fungi to herbicide toxicity. Felkner, I. C., and B. E. Worthy. 1990. Microbiological detection and quantitation of bioactive compounds in vegetation matrix by laser light scattering. Fresenius' Journal of Analytical Chemistry 338: 489-94. Fletcher, K., and B. Freedman. 1985. Effects of the herbicides glyphosate, 2,4,5trichlorophenoxyacetic acid, and 2,4-dichlorophenoxyacetic acid on forest litter decomposition. Canadian Journal of Forest Research 16: 6-9. Laboratory studies with two leaf litter and one forest floor substrate showed that the herbicides 2,4-dichlorophenoxyacetic acid (2,4-D),2,4,5-trichlorophenoxyacetic acid(2,4,5-T), a 50:50 mixture of these, and glyphosate all had toxic thresholds at which they reduced decomposition. However, in all cases, the thresholds were >50 times higher than residue concentrations that occur in the field after silvicultural herbicide treatments. In a field study at one site, no measurable 1-year postspray effects on litter decomposition were found among treatment plots sprayed at 0.0, 3.4 or 6.7 kg 2,4,5-T/ha. Garabito, G. E. A., Y. Sandoval, and A. H. Trujillo. 1991. Effect of some herbicidal compounds used in tropic areas on the microbial populations in soil. Reviews of Latinoam Microbiology 33: 191-201. The effects that some herbicidal compounds have on soil microbial populations were studied. The agrochemicals studied were: Glyphosate, Paraquat, Ametrine, 2,4-D, Atrazine and Atrazine + Terbutrine. The effect was measured by plate counts of bacteria, actinomycetes and fungi on three types of soils (Feozem, Castanozem and Chernozem) from Tecoman Col. Mexico, during nine weeks after treatment with agrochemicals in both field and laboratory conditions. Results showed that microbial populations tend to normality around 6 weeks after the exposure to agrochemicals. (Mexico). Ghassemi, M., S. Quinlivan, and M. Dellarco. 1982. Environmental effects of new herbicides for vegetation control in forestry. Environment International 7: 389-402. Environmental fate of and toxicity data for ammonium ethyl carbamoylphosphonate [fosamine ammonium), N-(phosphonomethyl)glycine [glyphosate], and 3-cyclohexyl-6-(dimethylamino)lmethyl1,3,5-triazine2,4-(lH,3H)dione(hexazinone) which appear promising for vegetation control in forestry are reviewed. None of the three herbicides is very persistent in soil, with halflives reported of about 7-10 days for fosamine ammonium, 8-19 weeks for glyphosate, and 1-16 months for hexazinone, depending on soil type and climatic conditions. Degradation in soil is primarily via microbial routes. Whereas hexazinone is a very mobile herbicide (with mobility dependent on soil type), fosamine ammonium and glyphosate are strongly adsorbed and not readily leached in many soil types. Based on bioassay results, the three herbicides exhibit "very low" to "low" toxicity. Fosamine ammonium is not mutagenic and was not teratogenic when fed to pregnant rats at 10,000 ppm. Data on hexazinone indicate no carcinogenic effect in rats, negative in bacterial and mammalian point mutation assays, and not embryotoxic or teratogenic at up to 5000 ppm in diet of rats. The three herbicides have minimumtonil effect on soil microorganisms and exhibit little or no potential for bioaccumulation. Limitations of the available data are discussed and suggestions are made for future studies. 101

25.

26.

27.

28.

Microflora and Fungi 29. Gomez, M. A., and M. A. Sagardoy. 1985. The effect of glyphosate on aerobic bacteria that colonize a sandy soil. Anales De Edafologia y Agrobiologia 44: 119-30. From a sandy soil of the semiarid zone of the province of Buenos Aires, Argentina, 519 bacterial strains were isolated and their genera were identified. Furthermore, the effect of the herbicide on the principal microbial groups was also studied. The results indicate that within the gram positive bacteria, coryneform bacteria and Bacillus spp. were predominant microorganisms. In the gram negative bacteria the Acinetobacter was found to be the dominant genus. The herbicide, at different doses, had little effect on the bacteria that colonize the soil. There was no deleterious effect detected on the potential of these bacterial strains as participants in the nitrogen cycle. The application of the herbicide to sterilized soil at the dose of 40 L/ha did not produce any deleterious effect on eurythermic bacteria of the types coryneform, Bacillus spp. or Acetobacter 307. However, after 10 days of incubation, the number of Acinetobacter 218 was found to decrease. The imperceptible effect produced by glyphosate on the bacteria which normally colonize such soil suggests that the use of the herbicide would not produce any significant alteration in that biological component of the ecosystem soil. (Argentina). . 1985. Influence of glyphosate herbicide on the microflora and mesofauna of a sandy soil in a semiarid region. Revista Latinoamericana De Microbiologia 27: 351-58. See Terrestrial Invertebrates Section. Grossbard, E. 1985. The Herbicide Glyphosate. Effects of glyphosate on the microflora: with reference to the decomposition of treated vegetation and interaction with some plant pathogens. E. Grossbard and D. Atkinson (editors). London: Butterworths. Pp. 159-185. (*) Although more research is required to clarify the interactions of glyphosate with the microflora a few common threads lead through the labyrinth of observations. In pure culture, many microbial species are inhibited by glyphosate. This effect is selective, variable in magnitude and frequently dose related. In terms of agricultural practice, inhibitory effects, especially of the cellulolytic fungi, occur at concentrations well above those normally used in agriculture. Moreover, the amount of herbicide which reaches the soil is smaller than that applied, because a certain proportion of glyphosate, which is foliar applied, is retained by the vegetation. Also glyphosate is absorbed and degraded in soil; one reason why the microorganisms examined tolerate, generally, the glyphosate in soil much better than in pure culture in laboratory media. Furthermore, the numbers of propagules of all microbial groups in soil increased following glyphosate treatment. Despite the shortcomings of techniques for enumerating microbial propagules in soil, this observation implies utilization by soil microflora of either glyphosate itself or its degradation products. A reassuring aspect of the behaviour of glyphosate is that it does not curtail nitrification in the soil, an activity affected adversely by some other herbicides. Mineralization of nitrogen, even at high concentrations of glyphosate, is enhanced, as is another important functioncellulose decomposition. On present evidence, however, it may be concluded that soil fertility per se is unlikely to be impaired by the agricultural use of glyphosate. (United Kingdom). Grossbard, E., and S. L. Cooper. 1974. The decay of cereal straw after spraying with paraquat and glyphosate. Proceedings of the British Weed Control Conference. 12:337-343. The effect of paraquat (1.7 kg/ha) and glyphosate (3.4 kg/ha) on the decay of C14-labelled rye leaves incubating on the soil surface and of unlabelled barley straw, covered by a shallow layer of soil, was studied in the lab. Paraquat and glyphosate had no effect on the decay of rye leaves on the surface and glyphosate did not affect degradation of barley straw. However, paraquat significantly suppressed the decay of buried barley straw as measured by weight loss, which was, after 16 weeks, between 40 and 55% of the initial weight for control and only between 18 and 26% with paraquat. The evolution of NH4NO3 increased the extent of decay of barley straw treated with paraquat by about 17% but at the end of the experiment the weight loss was still smaller than for control straw. (United Kingdom). Grossbard, E., and H. A. Davies. 1976. Specific microbial responses to herbicides. Weed Research 16: 163-69. Herbicides can exert adverse effects on the soil microflora, depending on concentration, though often at higher rates than the herbicidal dose. Examples refer to changes in microbial growth 102

30.

31.

32.

33.

Microflora and Fungi (asulam, linuron, paraquat) and equilibrium (metoxuron), respiration and nitrification in samples from field experiments (linuron, simazine) and laboratory experiments (bentazone, glyphosate, barban, etc), enzyme activities (urease, phosphatase, barban, chlorpropham, linuron) and the decay of sprayed vegetation (glyphosate, paraquat). (United Kingdom). 34. Grossbard, E., and D. Harris. 1977. Selective action of Gramoxone W and Roundup on Chaetomium globosum in relation to straw decay. Transactions of the British Mycological Society 69: 141-46. (United Kingdom). . 1981. Effects on straw decay of the herbicides paraquat and glyphosate in combination with nitrogen amendments. Annals of Applied Biology 98: 277-88. The effects of paraquat were studied at 4-7 kg a.i./ha and of glyphosate at 9.4 kg a.i./ha on the weight loss and rate of 02 uptake by stems of barley straw, incubated on the soil surface for various periods of time, and stems and leaves of wheat straw on the soil surface or covered with a shallow layer of soil. In general, paraquat reduced weight loss and frequently 02 uptake. Glyphosate normally had no effect but occasionally it was stimulatory. However, in isolated instances, paraquat was inactive and glyphosate slightly inhibitory. Leaves decayed faster than stems. Burial as well as the addition to straw of leaf protein, nitrochalk, urea and clover plants increased decomposition in all treatments, but nitrogen amendments were less effective in those with paraquat. For example, urea enhanced weight loss after 21 weeks from 19 to 34% of the initial weight in unsprayed wheat straw, from 12 to 24% in paraquat-treated and from 20 to 37% in glyphosate-treated straw, indicating that conditions stimulating decomposition may attenuate but not entirely eliminate the inhibitory effects of paraquat. A repeated addition of nitrogen to straw in various stages of decay increased 02 uptake in all treatments but least in the paraquat-treated straw, implying that a complete recovery from the initial curtailment of microbial activity, brought about by this herbicide, did not occur. Although the concentration of paraquat used was 9 times greater than the minimum recommended for field application, the degree of inhibition observed in these laboratory experiments was probably not sufficiently great to be of marked economic importance. (United Kingdom). Grossbard, E., and G. Wingfield. 1978. Effects of paraquat, aminotriazole and glyphosate on cellulose decomposition. Weed Research 18: 347-53. Effects of paraquat, activated aminotriazole and glyphosate on cellulose decomposition were studied in laboratory experiments. Herbicides were either incorporated in soil or sprayed onto cellulosic substrates which were incubated either on the surface of soil or buried. Paraquat applied directly to the substrates inhibited their decay, but not when applied to the soil. Aminotriazole incorporated in soil at 500 ppm strongly suppressed decomposition of buried calico, but effects varied with surface incubation. The only effects observed at 150 ppm were occasional stimulation of the decay of surface-incubated calico. Glyphosate applied either to soil or substrate usually stimulated cellulose decay. Numbers of propagules of cellulolytic fungi were decreased by aminotriazole and glyphosate applied to the soil and by glyphosate and paraquat applied to cellulose, but increased in soil treated with paraquat. In pure culture paraquat was the most, and aminotriazole the least, toxic to the 5 test fungi. Numbers of cellulolytic bacteria were decreased in soil treated with paraquat but the other 2 herbicides either did not affect or stimulated this group. (United Kingdom). Haahtela, K., S. Kilpi, and K. Kari. 1988. Effects of phenoxy acid herbicides and glyphosate on nitrogenase activity acetylene reduction in root-associated Azospirillum enterobacter and Klebsiella. FEMS (Federation of European Microbiological Societies) Microbiology Ecology 53: 123-28. Nitrogenase activity (C2H2 reduction) in root-associated Azospirillum lipoferum, Klebsiella pneumoniae, Enterobacter agglomerans and Pseudomonas sp. isolated from roots of Finnish grasses was assayed in the presence of glyphosate, the phenoxy acid herbicides 2-methyl-4-chlorophenoxy acetic acid (MCPA), 2,4-dichlorophenoxy acetic acid (2,4-D), (.+-.)-2-(2-methyl-4chlorophenoxy)propionic acid (Mecoprop) and (.+-.)-2-(2,4-dichlorophenoxy)propionic acid (Dichlorprop), and the commercial products Roundup, Nurmikko-hedonal, Mepro, and Dipro. In the presence of the phenoxy acid herbicides the nitrogenase activity of K. pneumoniae was significantly

35.

36.

37.

103

Microflora and Fungi inhibited, but that of E. agglomerans was stimulated. With the exception of Mepro and Mecroprop no phenoxy acid herbicides inhibited the nitrogenase activity of A. lipoferum and none that of Pseudomonas sp. Nurmikko-hedonal considerably stimulated the nitrogenase activity of E. agglomerans, and Pseudomonas sp. on the other hand, the nitrogenase activity of both K. pneumoniae and E. agglomerans was considerably repressed by glyphosate and Roundup, which also inhibited the growth of the bacteria. These chemicals had no effect on the growth of A. lipoferum and Pseudomonas sp., but stimulated their nitrogenase activity. (Finland). 38. Hallas, L. E., W. J. Adams, and M. A. Heitkamp. 1992. Glyphosate degradation by immobilized bacteria. Field studies with industrial wastewater effluent. Applied Environmental Microbiology 58: 1215-19. Immobilized bacteria have been shown in the laboratory to effectively remove glyphosate from wastewater effluent discharged from an activated sludge treatment system. Bacteria consortia in lab columns maintained a 99% glyphosate-degrading activity (GDA) at a hydraulic residence time of <20 min. In this study, a pilot plant (capacity, 45 liters/min) was used for a field demonstration. Initially, activated sludge was enriched for microbes with GDA during a 3-week biocarrier activation period. Wastewater effluent was then spiked with glyphosate and NH4Cl and recycled through the pilot plant column during start-up. Microbes with GDA were enhanced by maintaining the pH at <8 and adding yeast extract (<10 mg/liter). Once the consortia were stabilized, the column capacity for glyphosate removal was determined in a 60-day continuous-flow study. Waste containing 50 mg of glyphosate per liter was pumped at increasing flow rates until a steady state was reached. A microbial GDA of >90% was achieved at a 10-min hydraulic residence time (144 hydraulic turnovers per day). Additional studies showed that microbes with GDA were recoverable within (i) 5 days of an acid shock and (ii) 3 days after a 21-day dormancy (low-flow, low-maintenance) mode. These results suggest that full-scale use of immobilized bacteria can be a cost-effective and dependable technique for the biotreatment of industrial wastewater. Hallas, L. E., M. A. Heitkamp, and W. J. Adams. 1991. Glyphosate biodegradation by immobilized bacteria II. Pilot plant biotreatment of glyphosate in wastewater. 91st General Meeting of the American Society for Microbiology. p. 278. Harris, D., and E. Grossbard. 1979. Effects of the herbicides Gramoxone W and Roundup on Septoria nodorum. Transactions of the British Mycological Society 73: 27-33. The action of Gramoxone W (a.i. paraquat) and Roundup (a.i. glyphosate) on vegetative growth, spore formation, germination and pathogenicity of spores of Septoria nodorum Berk. to detached wheat leaf segments, was examined. Both herbicides had inhibitory effects, but Roundup was the more active. At 80 ppm it reduced the growth rates of four isolates by 50-55%, whilst Gramoxone W at 80 ppm produced a 25-40% reduction. Roundup at 80 ppm reduced spore formation by 90% but Gramoxone W was ineffective. Pretreatment with Roundup reduced spore germination by 60% even when the herbicide was absent from the germination medium. However, with Gramoxon W this inhibition occurred only when the herbicide was incorporated in the germination medium. A 14 potential carry-over of the active ingredient of Gramoxone W was indicated by the presence of C in 14washed spore suspensions prepared from cultures grown on media containing C labelled paraquat, and by the greater severity of lesions on detached leaves infected by spores from cultures grown with Gramoxone W, when compared to lesions produced by untreated spores. Roundup pretreated spores produced lesions less severe than control, except at high inoculum density, where in some experiments lesion severity was enhanced late in the incubation period. (United Kingdom). Harrison, R. D., and W. A. Gardner. 1992. Fungistasis of Beauveria bassiana by selected herbicides in soil. Journal of Entomological Science 27: 233-38. Oatmeal-dodine agar was used as a selective medium to enumerate colony-forming units (CFU's) of the entomogenous fungus Beauveria bassiana (Balsamo) Vuillemin in sterilized soil following the simultaneous application of conidia with six selected herbicides used in pecan management. At 3 days after treatment, significantly (P < 0.05) fewer CFU's were recovered from soils treated with recommended rates of either diuron, norflurazon, paraquat, simazine, or terbacil than were recovered from soils that were not treated with herbicides or were treated with glyphosate. By 7 days

39.

40.

41.

104

Microflora and Fungi after treatment, numbers of CFU's recovered from the soils did not differ significantly among the herbicide treatments and the controls. 42. Heinonen-Tanski, H., L. Montonen, L. R. Ervio, S. Junnila, and B. Pessala. 1985. The effects of chlorsulfuron, glyphosate, metribuzin and TCA on soil nitrification capacity and dehydrogenase activity. In: Comportement et Effects Secondaires des Pesticides dans le Sol. pp. 183-189. In field trials at Jokioinen and Ruukki, Finland, potatoes were treated with glyphosate, metribuzin or TCA and barley with chlorsulfuron all applied at the normal rate or 3 times this rate. The effects of the herbicides on soil dehydrogenase and nitrification activity were studied. Chlorsulfuron and glyphosate had little effect on soil dehydrogenase and nitrification, but metribuzin and TCA reduced both activities especially in clay soils. The effects of TCA were still evident 1 year after application. (Finland). Heinonen-Tanski, H., C. Rosenberg, H. Siltanen, S. Kilpi, and P. Simojoki. 1985. The effect of the annual use of pesticides on soil microorganisms, pesticide residues in the soil and barley yields. Pesticide Science 16: 341-48. A study has been made of the influence of pesticides used annually on soil microorganisms and crop yields (Finland). The persistence of these pesticides in the soil was also investigated. The herbicides MCPA, glyphosate, maleic hydrazide and triallate, and the insecticide parathion, were applied on experimental plots on which barley was grown during the years 1973-1981. The fungicide 2methoxyethylmercury chloride was used every year for dressing the seeds grown in pesticide-treated plots. The pesticide treatments did not affect significantly the numbers of several groups of soil microorganisms. A slight increase was, however, observed in the nitrification activity in the soil. The barley yields were on average higher on pesticide-treated plots than on controls because of successful weed control. Pesticide residues in the soil were generally very low; for example, for parathion they were below 0.02 mg/kg within 11 days, and for MCPA 0.06 mg/kg within 7 days. However, the glyphosate residue was 1.6 mg/kg in the autumn 2 days after the treatment, and the residue settled to a level of 0.2 mg/kg during the following summer. No clear dependence was observed between the residue level and the time between treatment and sampling. (Finland). Heinonen-Tanski, H., H. Siltanen, S. Kilpi, P. Simojoki, C. Rosenberg, and S. Makenen. 1986. The effect of the annual use of some pesticides on soil microorganisms, pesticide residues in soil and carrot (Daucus carota) yields. Pesticide Science 17: 135-42. The study deals with the effect of common, annually-used pesticides on soil microorganisms, pesticide residues in soil, and carrot (Daucus carota) yields in central Finland. Linuron residues in carrot roots were also analyzed. Thiram + lindane and dimethoate were applied from 1973-1981 at the commercially recommended doses on experimental plots of carrots, linuron was applied at twice the recommended rate from 1973-1979 and at the normal rate thereafter and in addition TCA was applied in 1978. Maleic hydrazide was used in the years 1973-1976, and glyphosate after 1977. The numbers of different soil microorganisms, their activities and the pesticide residues were studied from autumn 1978 to 1981. The pesticide treatments reduced the growth of soil algae but increased the total number of microorganisms and the number of aerobic sporeforming bacteria. Linuron residues in the soil were 0.9-2.8 mg/kg in the growing season and 1.2-1.7 mg/kg in the autumn, 3 months after application. The residues of glyphosate in the soil were 0.7 mg/kg in the autumn, 41 days after the treatment, and had declined to a level of about 0.2 mg/kg by the following summer. In the pesticidetreated plots the carrot yield was only 20-60% of the yield in the hand-weeded plots. The herbicide program controlled most of the annual weeds but not couchgrass Elymus repens and milk sow-thistle Sonchus arvensis. (Finland). Heitkamp, M. A., W. J. Adams, and L. E. Hallas. 1992. Glyphosate degradation by immobilized bacteria. Laboratory studies showing feasibility for glyphosate removal from waste water. Canadian Journal of Microbiology 38: 921-28. See Water Quality Section.

43.

44.

45.

105

Microflora and Fungi 46. Heitkamp, M. A., L. E. Hallas, and W. J. Adams. 1991. Glyphosate biodegradation by immobilized bacteria I. Laboratory studies showing treatment feasibility for aqueous waste streams. 91st General Meeting of the American Society for Microbiology. p. 278. Helweg, A. 1986. Side effects caused by pesticide combinations. FEMS (Federation of European Microbiological Societies) Symposium No. 33. Microbial Communities in Soil.V. Jensen, A. Kjoller, and L. H. Sorensen (editors). (Denmark). Hendricks, C. W., and A. N. Rhodes. 1992. Effect of glyphosate and nitrapyrin on selected bacterial populations in continuous-flow culture. Bulletin of Environmental Contamination and Toxicology 49: 417-24. This study was designed to use the continuous-flow method to determine the response of nitrifying and heterotrophic bacterial populations to treatments of a known nitrification inhibitor (nitrapyrin) and the widely used post-emergence herbicide, glyphosate. Inhibition of nitrification was dose-dependent. The values observed were 96.8, 100, and 100% for nitrapyrin and 26.8, 55.9, and 100% for glyphosate. Both chemicals did inhibit nitrifying bacteria in Amity soil, but their numbers were not appreciably changed. The sensitivity to nitrification inhibitors varied greatly among the genera of nitrifying bacteria. The increase in heterotrophic bacteria was probably the result of organic carbon into the system in the form of two treatment chemicals. Nitrite oxidizer numbers decreased in the treated soils. Since nitrapyrin inhibits ammonium oxidation specifically, it is likely that the decrease in nitrite oxidizers was the result of substrate loss. In glyphosate-treated soils the cause is not known and probably not related to nutrient limitation brought about by increased heterotrophic competition for nitrogen sources because of the high levels of ammonium added in the culture medium. The continuous-flow method to culture nitrifying bacteria has proven to be a viable alternative to the traditional methods to culture these bacteria. In this study, the continuous-flow system was used for both nitrification studies and for studies to determine the impact of chemical compounds on a soil process. Hendrix, P. F., and R. W. Parmelee. 1985. Decomposition nutrient loss and microarthropod densities in herbicide-treated grass litter in a Georgia piedmont USA agroecosystem. Soil Biology and Biochemistry 17: 421-28. A litterbag experiment was made to investigate the influence of the herbicides atrazine, paraquat and glyphosate on several aspects of the decomposition of grass litter in a fallow field. Litterbags containing dried Johnson grass (Sorghum halepense) leaves were dipped in herbicide solutions at recommended and 10 times recommended field application rates, placed in the field, and randomly collected every 3 weeks from July-November, 1982. At the highest treatment levels of paraquat and glyphosate, the leaves showed slower weight loss, faster losses of P, Ca and Mg, and higher densities of microfloral-grazing microarthropods than untreated controls. Using a conceptual model of the decomposition subsystem, it is hypothesized that herbicide treatment altered the system by promoting microbial utilization of the herbicide or additive as a carbon source; increasing the importance of microarthropod grazing relative to comminution; eliminating or reducing the importance of the predatory microarthropods; and increasing the rate of nutrient loss from the litter via microbial and microarthropod activity. The system thus became simplified with fewer recycling loops, accelerated soluble nutrient loss, and slower decay of carbon from the leaf tissue. Huang, J. W. 1993. Influence of herbicides on the growth of garden pea seedlings and their root disease pathogens. Plant Protection Bulletin (Taichung) 35 , no. 3 : 163-75 . Pre-emergence application of the herbicides, alachlor (2-chloro-2',6'-diethyl -N- (methoxymethyl) acetanilide), butachlor (chloro-2',6'- diethyl -N- (butoxy methyl acetanilde)), stomp (N-(1-ethyl propyl) -2,6-dinitro -3,4-xylidine), glyphosate (N- (phosphonomethyl) glycine), trifluralin (trifluoro -2,6dinitro -N,N- dipropyl -P- toluidine), paraquat (1,1'- dimethyl -4,4'- bipyridylium dimethyl sulfate) and atrazine (2-chloro -4- isopropylamino ethylamino -s-triazine), to soil was associated with injury symptoms of garden pea seedlings (Pisum sativum cv. Taichung No.11). Seedlings grown in soil 106

47.

48.

49.

50.

Microflora and Fungi treated with alachlor or butachlor exhibited basal stem and root necrosis, yellowing by atrazine, and stunting by stomp or trifluralin. In addition, the formation of root nodules was significantly enhanced by stomp, but suppressed by glyphosate. Microbial population in the soil varied with the kind of herbicides used. Actinomycetous and bacterial populations were increased significantly by paraquat and alachor, respectively, two weeks after treatment. However, a higher level of bacterial population was maintained for 4 weeks only in trifluralin- and glyphosate-treated soils. Although alachlor, stomp, and trifluralin also increased soil fungal population two weeks after treatment, their increases were small compared with the others including no treatment control. Radial growth of F. solani f. sp. pisi and R. solani AG-4 was assessed by incorporating each one of seven herbicides at concentrations ranging from 0 to 5000 ppm within 2% water agar. Radial growth of the fungi was significantly reduced with increment of herbicide dosage. Furthermore, F. solani f. sp. pisi was able to tolerate much more toxicity of herbicides than R. solani AG-4. In greenhouse tests, treatment of soil with herbicides, alachlor, butachlor, stomp, glyphosate, and trifluralin at a rate of 6.25 kg/ha, increased Rhizoctonia blight of garden pea seedlings caused by R. solani AG-4 by 20 to 40%. Alachor, butachlor at 6.25 kg/ha, and atrazine at 0.625 kg/ha also predisposed 65%, 19% and 37% of garden peas, respectively, to infection by the root rot pathogen F. solani f. sp. pisi. (Taiwan) . 51. Ismail, B. S., A. J. Kader, and O. Omar. 1995. Effects of glyphosate on cellulose decomposition in two soils. Folia Microbiologica 40, no. 5: 499-502. See Plant and Soil Residues Section. Jacob, G. S., J. R. Garbow, L. E. Hallas, N. M. Kimack, G. M. Kishore, and J. Schaefer. 1988. Metabolism of glyphosate in Pseudomonas sp. strain LBR. Applied and Environmental Microbiology 54: 2953-58. Metabolism of glyphosate (N-phosphonomethylglycine) by Pseudomonas sp. strain LBR, a bacterium isolated from a glyphosate process waste stream, was examined by a combination of solidstate C13 nuclear magnetic resonance experiments and analysis of the phosphanate composition of the growth medium. Pseudomonas sp. strain LBR was capable of eliminating 20 mM glyphosate from the growth medium, an amount approximately 20-fold greater than that reported for any other microorganism to date. The bacterium degraded high levels of glyphosate, primarily by converting it to aminomethylphosphate, followed by release into the growth medium. Only a small amount of aminomethylphosphonate (about 0.5 to 0.7 mM), which is needed to supply phosphorus for growth, could be metabolized by the microorganism. Solid-state C13 nuclear magnetic resonance analysis of strain LBR grown on 1 mM (C13-2, N15)glyphosate showed that about 5% of the glyphosate was degraded by a separate pathway involving breakdown of glyphosate to glycine, a pathway first observed in Pseudomonas sp. strain PG2982. Thus, Pseudomonas sp. strain LBR appears to possess two distinct routes for glyphosate detoxification. (United Kingdom). Junnila, S. 1986. An overview of trials concerning mobility and transport of herbicides. Nordisk Jordbrugsforskning 68: 384-85. This paper was presented at the Nordiske Jordbrugsforskeres Foreming Workshop on leaching of chemical control compounds, held in Uppsala in Nov. 1985. In field trials at the Agricultural Research Centre, Jokioinen, and at Norra-Osterbotten starting in 1982, the persistence and mobility of 6 herbicides and their effect on soil microorganisms were investigated. Glyphosate was applied in autumn and TCA in spring and autumn before potatoes were planted. Metribuzin and flurochloridone were applied to potatoes in June. In spring cereals, chlorsulfuron and metsulfuron-methyl were tested at 4 (normal rate) and 12 g/ha. Soil samples were taken at 0-5, 5-15 and 15-25 cm depths, 1 day and 1 month after spraying, and also at the beginning and end of the growth period, so long as test plants in greenhouses showed that residues were still present. Trials were on sand, loam and peat soils at Jokioinen and on sand and peat soils at Norra-Osterbotten. In sandy soils TCA moved quickly to a depth of 25 cm and disappeared after 6 months. In loam, TCA was more readily bound, and residues were found after a year at the normal rate and after 2.5 years at the higher rate. TCA and chlorsulfuron tended to move vertically through the soil. With spring application, TCA reached a depth of 25 cm after 1 month. Glyphosate was strongly adsorbed by soil. Ryegrass (Lolium sp.) test plants showed damage 6 months after glyphosate had been applied to the loam and sand soils and 1 year after application to peat soils in 1982. After spraying in 1983, plants were still damaged 1.5 years later. 107

52.

53.

Microflora and Fungi Metribuzin residues were found more than or equal to 1.5 years after application at 0.75 kg/ha and 2 years after application at 1.4 kg/ha. In 1983 metribuzin reached a depth of 25 cm in less than a month in sand and peat and by autumn in loam. In 1984, when rain during spraying was heavy, it reached a depth of 15 cm in 1 day. Flurochlorindone was found at a depth of 25 cm 1 day after sowing after abundant rain in 1984. At normal rates, residues remained for at least 1 growing season. Chlorosulfuron and metsufuron-methyl were both adsorbed quickly by all soil types and residues remained for at least one growing season but often for 1 year, especially at 3 times the normal rate. Soil tended to adsorb metsulfuron-methyl more than chlorsulfuron. (Sweden). 54. Kassaby, F. Y., and G. Hepworth. 1987. Phytophthora cinnamomi effects of herbicides on radial growth sporangial production inoculum potential and root disease in Pinus radiata. Soil Biology and Biochemistry 19: 437-42. An epidemic of severe root rot of Pinus radiata seedlings caused by Phytophthora cinnamomi in the 1970's at Benalla nursery, 220 km north-east of Melbourne (Victoria) was followed by a rapid disappearance of the disease in 1981. This phenomenon coincided with the application of simazine and propazine for weed control between 1971 and 1980 and glyphosate and chlorthal dimethyl after 1980. In laboratory and greenhouse tests, chlorthal dimethyl and glyphosate significantly (P < 0.01) reduced radial growth, sporangial production and inoculum potential of P. cinnamomi, whereas propazine and simazine exhibited mild fungitoxic properties (in vitro), and had a stimulatory effect on the production of sporangia and inoculum potential of the pathogen. (Australia). Knowles, W. S., K. S. Anderson, S. S. Andrew, D. P. Phillion, J. E. Ream, K. A. Johnson, and J. A. Sikorski. 1993. Synthesis and characterization of N-amino-glyphosate as a potent analog inhibitor of E. coli EPSP synthase. Bioorganic and Medicinal Chemistry Letters 3 , no. 12 : 2863-68 . All previous attempts to identify glyphosate analogs which retain their potency against the known biological target, EPSP synthase, have been unsuccessful. Consequently, the glyphosate binding site was thought to be extremely specific in this system. Here we report the novel N-amino glyphosate analog 3 as the first successful modification of the glyphosate skeleton which exhibits inhibitor properties comparable to glyphosate. Korol, R. V. 1985. Regulation of glyphosate content in the soil. Gigiena I Sanitariya 8: 71-72. (Soviet Union). Kruglov, Y. V., N. B. Gersh, and M. Shtal'berg. 1980. The influence of glyphosate on the soil microflora. Khimiya v Sel'Skom Khozayistive 18: 42-44. (Soviet Union) . Levesque, C. A., K. Beckenbach, D. L. Baillie, and J. E. Rahe. 1993. Pathogenicity and DNA restriction fragment length polymorphism of isolates of Pythium spp. from glyphosate-treated seedlings. Mycological Research 97, no. 3 : 307-12 . Sixty Pythium isolates from glyphosate-treated wheat or green bush bean seedlings grown in mineral or organic soils were examined with regard to possible host and site (soil type) specificity. Pythium ultimum dominated the Pythium population isolated from glyphosate-treated root systems planted in organic soil (20 out of 27 isolates), whereas P. sylvaticum was slightly predominant in mineral soil (18 out of 33 isolates). Isolates were compared on the basis of differences in DNA sequences detected by size separation after restriction endonuclease digestion of total DNA. A computational technique to calculate distance matrix from RFLP data was developed. Intraspecific DNA variation did not correlate with host from which the isolates were obtained and there was a correlation between the DNA-based groupings of P. ultimum isolates and the soil type in which the colonized glyphosate-treated seedlings were grown. Two isolates from each fungal species were selected for evaluation of in vivo pathogenicity and host specificity. All four isolates reduced seedling emergence in wheat but only P. ultimum caused a reduction of emergence in beans. There was no indication of host specialization among the isolates tested, i.e. P. sylvaticum or P. ultimum isolates obtained from wheat did not cause a higher incidence of seedling blight of wheat than of beans and vice versa.

55.

56. 57.

58.

108

Microflora and Fungi 59. Levesque, C. A., and J. A. Rahe. 1992. Herbicide interactions with fungal root pathogens with special reference to glyphosate. In: Annual Review of Phytopathology. R.J. Cook (editor) Annual Reviews Inc., Palo Alto, California, USA. Vol. 30. pp. 579-602. Levesque, C. A., and J. E. Rahe. 1989. Fungal colonization of glyphosate-treated plants. Canadian Journal of Plant Pathology 11: 193-94. The sensitivity of plants to the herbicide glyphosate was reduced significantly by heat treatment of soil in all plant species tested so far (3 monocots and 4 dicots). The efficacy of glyphosate on wheat or beans grown in heat-treated soil was restored when untreated aqueous extracts of soil were added to the heat-treated soil. Fungal colonization of roots of wheat and bean seedlings was detected within 24 hours after glyphosate treatment. Pythium and Fusarium spp. were the predominant colonizers of glyphosate treated plants. In wet loam soil (field capacity) at 17 oC or 25oC, the rate of root colonization by Pythium spp. was twice that of Fusarium spp. Rate of colonization was measured as the number of colony forming units in the soil system over time. In dry soil (water suction > 1.0 Bar) Pythium and Fusarium spp. had similar colonization rates. Emergency of wheat and bean seedlings in autoclaved soil inoculated with Pythium spp. isolated from glyphosate-treated plants was drastically reduced (2 to 30 times) compared to that from seeds germinating in control soil. Levesque, C. A., J. E. Rahe, and D. M. Eaves. 1987. Effects of glyphosate on Fusarium spp.: its influence on root colonization of weeds, propagule density in the soil, and crop emergence. Canadian Journal of Microbiology 33: 354-60. Glyphosate is a broad spectrum herbicide that can lead to root rot like damage on crops. This study was undertaken to investigate the effect of glyphosate on the root-colonizing Fusarium spp. The research was conducted at two sites. Site one was densely covered with perennial weeds, and site two with annuals. At site one, spraying the weed cover with glyphosate increased (p < 0.05) the level of colonization by Fusarium spp. in Ranunculus repens and Holcus lanatus, but not in Stellaria media and Plantago lanceolata. At site two, glyphosate enhanced colonization in Sperqula arvensis, Stellaria media, Echinochloa crusgalli, and Chenopodium albuim, but not in Capsella bursa-pastoris and Polygonum persicaria. At both sites, the number of colony-forming units of Fusarium spp. per gram of dried soil was increased by the application of glyphosate. Nevertheless, crops subsequently sown in the field containing the annual weeds were not detrimentally affected by glyphosate treatment of these weeds. . 1992. The effect of soil heat treatment and microflora on the efficacy of glyphosate in seedlings. Weed Research 32: 363-73. Seedlings of wheat (Triticum aestivum L.) and beans (Phaseolus vulgaris L.) were less sensitive to glyphosate when grown in heat-treated soil than in raw soil. Pythium spp. and Fusarium spp. were not detected in heat-treated loam or muck soils at the time of glyphosate treatment, although fungi of several other genera were present. The efficacy of glyphosate on wheat or beans grown in heat-treated loam soil was restored when untreated aqueous soil extracts were added to the heattreated soil. Bean seedlings grown in five different soil types varied in their sensitivity to glyphosate. The variation in LD50 among autoclaved soils was lower than that among raw soils. Between 13- to 47-fold more glyphosate was required to kill the bean seedlings in any of the autoclaved soils compared with their corresponding raw soils. This differential effect was not observed on bean seedlings sprayed with either 2,4-D or paraquat. LD50 values for glyphosate on apple (Malus domestica Borkh.) seedlings growing in previously sterilized loam soil were reduced by inoculation of the soil with a representative Fusarium sp. or Pythium sp. obtained earlier from apple seedlings treated with glyphosate, but not by a Cylindrocarpon sp. from apple or Pythium ultimum Trow from glyphosatetreated bean. The efficacy of glyphosate on bean, wheat or apple seedlings can be affected by changes in certain microbial components of the soil. . 1993. Fungal colonization of glyphosate-treated seedlings using a new root plating technique. Mycological Research 97, no. 3 : 299-306 . A new technique for assessing the number and location of fungal colonizers of entire root systems was developed, and the effect of the herbicide glyphosate on fungal colonization of roots evaluated. Fungal colonization of roots of wheat and green bush bean seedlings grown at a 25:18 109

60.

61.

62.

63.

Microflora and Fungi degree C day:night regime took place less than 48 h after treatment of the plants with glyphosate. For both plant species grown under each of four environmental conditions (a combination of two temperatures, 17 degree and 25 degree, and two soil matric potentials, -6 and -100 kPa), Pythium spp. were the most frequent colonizers of glyphosate-treated seedlings and Fusarium spp. were the secondmost frequent colonizers. Colonization of control seedlings by Pythium spp. was only observed in beans grown at 17 degree. In glyphosate-treated wheat seedlings, less colonization by Pythium spp. occurred at 17 degree than at 25 degree, but soil water content had no significant effect. Under low soil water content, colonization by Fusarium spp. was always higher in glyphosate-treated bean or wheat seedlings than in control seedlings, whereas this differential effect was observed only for wheat grown at 25 degree under higher soil moisture. 64. Liu, C. M., P. A. McLean, C. C. Sookdeo, and F. C. Cannon. 1991. Degradation of the herbicide glyphosate by members of the family Rhizobiaceae. Applied Environmental Microbiology 57: 1799-804. Several strains of the family Rhizobiaceae were tested for their ability to degrade the phosphonate herbicide glyphosate (isopropylamine salt of N-phosphonomethylglycine). All organisms tested (seven Rhizobium meliloti strains, Rhizobium leguminosarum, Rhizobium galega, Rhizobium trifolii, Agrobacterium rhizogenes, and Agrobacterium tumefaciens) were able to grow on glyphosate as the sole source of phosphorus in the presence of the aromatic amino acids, although growth on glyphosate was not as fast as on Pi. These results suggest that glyphosate degradation ability is widespread in the family Rhizobiaceae. Uptake and metabolism of glyphosate were studied by using R. meliloti 1021. Sarcosine was found to be the immediate breakdown product, indicating that the initial cleavage of glyphosate was at the C.sbd.P bond. Therefore, glyphosate breakdown in R. meliloti 1021 is achieved by a C.sbd.P lyase activity. Liu, L., Z. K. Punja, and J. E. Rahe. 1992. Phytoalexin production in bean roots grown on sterile media and in natural soil. Phytopathology 82: 1085. Application of the herbicide glyphosate at sublethal doses has been previously shown to enhance infection by Pythium and Fusarium spp. on bean roots, and reduced the levels of phaseollin, phaseollinisoflavan, phaseollidin and kievitone in bean leaves. The purpose of this research was to determine if glyphosate influences phytoalexin production in roots, which in turn could influence root susceptibility to fungal infection. Production of phytoalexins in healthy roots grown in sterile growth media including 0.65% water agar, silica sand and metro-mix was compared to that in natural soil. Seedlings were grown for 7-10 days, and the phytoalexins were extracted in 95% boiling ETOH, and quantitative analysis was performed by HPLC. Only a trace amount of phaseollin was found in roots grown on water agar. Large amounts of phaseollinisoflavan and phaseollin were produced in roots grown in metro-mix and natural soil. Both phytoalexins were also found in roots grown on silica sand, but at a lower concentration. The effects of glyphosate on altering the concentration or ratios of phytoalexins produced in bean roots are being investigated. . 1995. Effect of Pythium spp. and glyphosate on phytoalexin production and exudation by bean (Phaseolus vulgaris L.) roots grown in different media. Physiological and Molecular Plant Pathology 47, no. 6 : 391-405. Kievitone, phaseollinisoflavan and phaseollin were detected in roots of bean seedlings (Phaseolus vulgaris L.) grown in natural soil. Comparison of phytoalexin production by roots grown in different media indicated that these phytoalexins were probably induced by microorganisms in soil. The influence of common root rot pathogens of bean, Pythium spp., on phytoalexin production was determined. Pythium ultimum elicited kievitone, phaseollinisoflavan and phaseollin in roots grown in sterilized silica sand. P. sylvaticum induced only kievitone and phaseollin in the same growth medium. Glyphosate did not significantly affect the accumulation of phytoalexins within 3 days. However, by day 5, significantly more phaseollin was detected in the roots of Pythium inoculated plants treated with glyphosate than in Pythium inoculated plants not treated with glyphosate. In a hydroponic system, both Pythium spp. elicited accumulation of kievitone and phaseollin in root tissue, and both phytoalexins were exuded into the bathing solution. Glyphosate application did not significantly affect accumulation or exudation of phytoalexins by bean roots in the hydroponic system. The results from this study

65.

66.

110

Microflora and Fungi illustrate the nature and extent of phytoalexin production by bean roots in the absence and presence of microbes. 67. Liu, Z. Q., and A. Fer. 1990. Effect of a parasite Cuscuta lupuliformis Krock on the redistribution of two systemic herbicides applied on a legume Phaseolus aureus Roxb. C R Academy of Science Series III Sci VIE 311: 333-38. The redistribution of two 14C-labelled herbicides is compared within a legume (Phaseolus aureus Roxb.) parasitized or not by dodder (Cuscuta lupuliformis Krock.). When pendimethalin, an ambimobile herbicide, is applied on the roots of Phaseolus, dodder has almost no influence on the redistribution-via the xylem-of the xenobiotic substance within the legume. In contrast, when pendimethalin is applied on the leaves of Phaseolus the occurrence of dodder induces a strong modification of the redistribution-via the phloem-of the herbicide. Also, when glyphosate, a typical phloem-mobile herbicide, is applied on the leaves, its redistribution within the legume is strongly modified by the occurrence of dodder. As a rule, dodder stimulates the exportation of phloem-mobile herbicides out of the host leaf and accumulates them. (France). Lonsjo, H., J. Stark, L. Torstensson, and B. Wessen. 1980. Glyphosate: decomposition and effects on biological processes in soil. Weeds and Weed Control, 21st Swedish Weed Conference, 1/2: 140-146. (Sweden). Majumder, K., A. Selvapandiyan, F. A. Fattah, N. Arora, S. Ahmad, and R. K. Bhatnagar. 1995. 5Enolpyruvylshikimate -3-phosphate synthase of Bacillus subtilis is an allosteric enzyme: Analysis of Arg24 fwdarw Asp, Pro105 fwdarw Ser and His385 fwdarw Lys mutations suggests a hidden phosphoenolpyruvate -binding site. European Journal of Biochemistry 229, no. 1: 99106 . 5-Enolpyruvylshikimate -3-phosphate synthase of Bacillus subtilis has been cloned, expressed and purified to near homogeneity. Clustal alignment of the amino acid sequences from different bacteria revealed several conserved residues located in the N-terminal, middle and C-terminal domains. The role of conserved Arg24, Pro105, and His385 residues has been examined by sitedirected mutagenesis. Steady-state kinetic analysis of the native synthase exhibited allosteric behaviour, a feature thought to be unique amongst bacterial and plant 5-enolpyruvylshikimate -3phosphate synthase enzymes investigated so far. Both substrates, phosphonenolpyruvate (P-pyruvate) and shikimate-3-phosphate have multiple interaction sites. There are two sites for P-pyruvate binding, catalytic and non-catalytic. Glyphosate (N-phosphonomethyl glycine) competes for binding at the catalytic site and does not interact at the secondary site. Glyphosate in the absence of ammonium ions increases cooperativity of P-pyruvate binding and favors dimerization of the enzyme through an interaction between P-pyruvate-binding sites. The ammonium-ion-activated 5-enolpyruvylshikimate -3phosphate synthase displays no cooperativity with respect to P-pyruvate. Absence of ammonium ions decreases affinity for substrates and introduces cooperativity. Cooperativity was also introduced in the enzyme by point mutations, Arg24-Asp and His385-Lys. The latter mutant of the native enzyme exists as a dimer and aggregates to a tetrameric form in the presence of glyphosate. The occurrence of multimeric forms of the synthase has been demonstrated by staining for the enzyme activity on the native gel and by resolving purified enzyme preparations on a sucrose density gradient. A model describing the alteration in the aggregation status of the enzyme by the inhibitor, activator and the substrates has been proposed. (India) . Mallik, M. A. B., and K. Tesfai. 1985. Pesticidal effect on soybean and Rhizobia symbiosis. Plant and Soil 85: 33-42. Relative compatibility of selected pesticides at 2 levels of application: recommended rate and 5 times or 10 times with soybean-rhizobia symbiosis was tested in pot culture experiments using a prepared peat inoculant. PCNB (pentachloronitro benzene), carboxin and carboxin + captan at recommended level were innocuous to growth, nodulation, N2-fixation and total nitrogen content of shoot. Carboxin and carboxin + captan but not PCNB at 10 times recommended level proved detrimental to nodulation and N2-fixation. Carbaryl and malathion at recommended level had no adverse effect but at 10 times recommended level severely reduced N2-fixation but not other 111

68.

69.

70.

Microflora and Fungi parameters. Acephate, diazinon and toxaphene at both levels reduced N2-fixation and total nitrogen content but not growth and nodulation. All five herbicides used at recommended and 5 times recommended level adversely affected nodulation and N2-fixation. Glyphosate proved least toxic to all parameters. 2,4-D at recommended level was less harmful to nodulation and N2-fixation than trifluralin, alachlor and metribuzin. 71. Marsh, J. A. P., H. A. Davies, and E. Grossbard. 1977. The effect of herbicides on respiration and transformation of nitrogen in 2 soils. Part 1 metribuzin and glyphosate. Weed Research 17: 7782. The effects of metribuzin and glyphosate at 100 ppm on microbial C02 evolution and nitrogen transformation in 2 soils were investigated in the laboratory. Both herbicides reduced CO2 evolution from Boddington Barn soil (organic carbon content 1.5%, pH 6.6) at some dates, but neither gave any consistent effects on Triangle soil (organic carbon content 4.0%, pH 5.1). Both metribuzin and glyphosate stimulated mineralization of nitrogen for at least 9 weeks. Only metribuzin on Triangle soil gave any indication of inhibition of nitrification. Metribuzin degraded more rapidly in Triangle soil than in Boddington Barn. (United Kingdom). Martensson, A. M. 1992. Assessing anthropogenic impact on heterotrophic nitrogen fixing soil microorganisms. Proceedings of the International Symposium on Environmental Aspects of Pesticide Microbiology J.P.E. Anderson et al. (editors). pp. 115-120. (Sweden). . 1992. Effects of agrochemicals and heavy metals on fast-growing Rhizobia and their symbiosis with small-seeded legumes. Soil Biology and Biochemistry 24: 435-45. The effect of potentially hazardous agrochemicals including fungicides, herbicides and heavy metals on symbiotic nitrogen fixation (in Trifolium pratense, Medicago sativa, Lotus corniculatus) have been investigated. The substances were tested with eight rhizobial strains from three cross-inoculation groups: Rhizobium leguminosarum b.v. trifolii, R. meliloti and R. loti in pure culture studies. Bacteria were obtained from a culture collection or from soils. Sensitivity of the bacteria to the agrochemicals and heavy metals varied. None of the bacteria were tolerant to all chemicals. No difference in tolerance between cross-inoculation groups existed. Bacteria were able to multiply at concentrations of agrochenmicals equal to or higher than recommended field-application rates. Heavy metals concentrations that severely inhibited growth were far lower than the highest amounts allowed under the current Commission of the European Communities' guidelines for environmental protection. Bacterial growth in presence of the agrochemicals and heavy metals, apart from glyphosate and zinc, did not influence nodulation ability of the strains. Development of uninoculated plants was inhibited at increasing concentrations of all compounds, red clover being most sensitive. Herbicides were most harmful, with injuries occurring at levels 1/10-1/10,000 of recommended applied concentrations. Uninoculated plants were less tolerant to agrochemicals, but were more tolerant to heavy metals compared to the bacteria. Root hair deformations similar to bacterial-induced root hair deformations were induced by bentazone, chlorsulphuron and monochlorophenoxyacetic acid on uninoculated plants. Symbiotic interactions were adversely affected by several of the agrochemicals. Bacterial-induced root hair deformations necessary for nodulation decreased in the presence of benomyl, bentazone, chlorsulphuron, fenpropimorph, mancozeb and monochlorophenoxyacetic acid. Fenpropimorph and mancozeb did not cause root hair deformations at increasing concentrations, indicating that these may inhibit nodulation under field conditions. Nodule development was inhibited at increased levels of bentazone, chlorsulphuron, glyphosate and mancozeb. Dry matter production of nodulated plants was adversely affected by bentazone and chlorsulphuron, indicating disturbances in nodule function. (Sweden). McAuliffe, K. S., L. E. Hallas, and C. F. Kulpa. 1990. Glyphosate degradation by Agrobacterium radiobacter isolated from activated sludge. Journal of Industrial Microbiology 6: 219-22. Two species of bacteria capable of growth on N-phosphonomethylglycine (glyphosate) were isolated from a bench scale sequencing batch reactor degrading a waste stream containing glyphosate. The enrichment and isolation medium contained defined salts and glyphosate as the sole carbon and

72.

73.

74.

112

Microflora and Fungi energy source. Glyphosate was stoichiometrically degraded to aminomethylphosphonic acid (AMPA). The bacteria have been identified as Agrobacterium radiobacter and Achromobacter Group V D. 75. Mekwatanakarn, P., and K. Sivasithamparam. 1987. Effect of certain herbicides on growth and pathogenicity of take-all fungus on wheat Triticum aestivum L. Biology and Fertility of Soils 5: 31-35. The effects of spray seed (Diquat + Paraquat), Roundup (glyphosate), Banvel-D (dicamba), Treflan (trifluralin), Glean (chlorsulfuron) and Dacthal (chlorthal dimethyl) at concentrations of 0-500 ppm product on the vegetative growth, vigour and pathogenicity of Gaeumannomyces graminis var. tritici (GGT) on wheat were examined. All herbicides with the exception of Dicamba and Chlorsulfuron inhibited fungal growth of potato dextrose agar (PDA) at concentrations 10-500-fold of rates recommended for use in the field. The vegetative growth of the pathogen growing out of straw colonized on PDA supplemented with 100 ppm Diquat + Paraquat or glyphosate was reduced by 47.4% and 42.4%, respectively. When portions of the colonies were subcultured onto unamended PDA, their growth and the pathogenicity of straw pieces colonized by these subcultures were found to be unaltered. Straw colonized by GGT on agar amended with concentrations of Diquat + Paraquat or at all concentrations of glyphosate produced less root disease in wheat seedlings in comparison to those colonized on unamended agar. It is proposed that the reduced pathogenicity of inocula prepared on agar amended with these two herbicides is due to poor colonization by the pathogen of straw on the media, and that a similar effect on saprophytic colonization in the field could lead to a reduction in the field inocula of the pathogen. (Australia). . 1987. Effect of certain herbicides on saprophytic survival and biological suppression of the take-all fungus. New Phycologist 106: 153-60. Saprophytic survival and pathogenicity of Gaeumannomyces graminis (Sacc.) ARX and Olivier var. tritici Walker were evaluated in sterile and unsterile western Australian wheat field soil treated with the herbicides glyphosate, Diquat + Paraquat, or Trifluralin. Survival in colonized straws and subsequent pathogenicity were not affected by the herbicide treatments in sterile soil. In unsterile soil, however, survival and pathogenicity were higher in glyphosate or Diquat + Paraquat-treated than in untreated or Trifluralin-treated soil. Incorporation of untreated natural soil into fumigated soil, to give concentrations of 1 and 10% natural soil in the mix, reduced disease in comparison with that in 100% sterile or 100% natural soil. Disease in fumigated soil increased progressively with increasing concentrations of glyphosate-treated, unsterile soil. Pre-treatment of wheat plants or soil with glyphosate before exposure of the host to the fungal pathogen or to other soil micro-organisms showed that the increase in disease following glyphosate treatment was not related to the direct effect of the herbicide on the host. (Australia). . 1987. Effect of certain herbicides on soil microbial populations and their influence on saprophytic growth in soil and pathogenicity of take-all fungus. Biology and Fertility of Soils 5: 175-80. The application of Diquat + Paraquat, glyphosate and Trifluralin to unsterilized field soil increased take-all caused by the fungus, Gaeumannomyces graminis var. tritici Walker by 13.0%, 16.6% and 10.8% respectively, while no effect on disease was recorded in sterilized soil treated with the same herbicides. The herbicides tested had no effect on the saprophytic growth of the pathogen with the exception of glyphosate, which increased its growth in unsterilized soil. The application of Diquat + Paraquat and glyphosate to unsterile soil had no effect on the numbers of Actinomycetes. The Diquat + Paraquat treatment, however, increased populations of fungi while the glyphosate decreased the numbers of bacteria. The proportion of soil fungi antagonistic to the pathogen was reduced in glyphosate-treated soil. The frequency of occurrence of Eupenicillium euglaucum (V. beyma) Stolk and Samson (strain B), and Penicillium verruculosum Peyr. (strain B), which were strong and low level antagonists of GGT on agar, were reduced in their occurrence in soil by 7.7% and 2.5% respectively, following glyphosate treatment. Moreover, the numbers of Aspergillus viridi-nutans Ducker and Thrower, which showed moderate antagonism to the pathogen, was decreased by 1.9% and 4.1% in Diquat + Paraquat and glyphosate treatments respectively. The proportion of antagonists rather than total numbers of fungi appears to be related to the treatment effect observed on the soil growth and pathogenicity of G. graminis var. tritici in our investigation. The increase in disease of 113

76.

77.

Microflora and Fungi wheat in certain herbicide-treated soils may be due to the shift in soil microbial populations away from those which are antagonistic to the pathogen. (Australia). 78. 79. Mercer, P. C., and T. W. Fraser. 1986. Microorganisms associated with the retting of flax treated with the herbicide glyphosate. Annals of Applied Biology 109: 509-21. Michailides, T. J., and R. A. Spotts. 1991. Effects of certain herbicides on the fate of sporangiospores of Mucor piriformis and conidia of Botrytis cinerea and Penicillium expansum. Pesticide Science 33: 11-22. The effects of several herbicides used in pome fruit orchards on the germination of spores and growth of mycelia of Botrytis cinerea, Mucor piriformis and Penicillium expansum in vitro and the survival of propagules of these fungi in soil were studied. Diuron in agar at 4-128 g/ml reduced germination of spores of B. cinerea and M. piriformis, and 2,4-D and paraquat at 32 g/ml similarly affected B. cinerea and P. expansum. Several herbicides at 128 g/ml in agar reduced growth of B. cinerea and M. piriformis but were ineffective against P. expansum. Propagule survival levels of the three fungi generally were lower in both autoclaved and non-autoclaved soil amended with herbicides than in non-amended soil. This effect was greatest in non-autoclaved soil, suggesting involvement of microbial antagonists. The most effective herbicides for reduction of fungal propagules in soil were 2,4-D, diuron, and paraquat. Moore, J. K., H. D. Braymer, and A. D. Larson. 1983. Isolation of a Pseudomonas sp. which utilizes the phosphonate herbicide glyphosate. Applied and Environmental Microbiology 46: 316-20. A strain of bacteria has been isolated which rapidly and efficiently utilizes the herbicide glyphosate (N-phosphonomethylglycine) as its sole phosphorus source in a synthetic medium. The strain (PG2982) was isolated by subculturing Pseudomonas aeruginosa ATTCC 9027 in a synthetic broth medium containing glyphosate as the sole phosphorus source. Strain PG2982 differs from the culture of P. aeruginosa in that it is nonflagellated, does not produce pyocyanin, and has an absolute requirement for thiamine. Strain PG2982 has been tentatively identified as a Pseudomonas sp. strain by its biochemical activities and moles percent guanine plus cytosine. Measurements of glyphosate with an amino acid analyzer show that glyphosate rapidly disappears from the medium during exponential growth of strain PG2982. In batch culture at 30oC, this isolate completely utilized 1.0 mM glyphosate in 96 h and yielded a cell density equal to that obtained with 1.0 mM phosphate as the phosphorus source. However, a longer lag phase and greater generation time were noted in the glyphosate-containing medium. Strain PG2982 can efficiently utilize glyphosate as an alternate phosphorus source. Moorman, T. B., J. M. Becceril, J. Lydon, S. O. Duke, and H. Matsumoto. 1990. Effects of the herbicide glyphosate on Bradyrhizobium japonicum growth inhibition and production of protocatechuate. 90th Annual Meeting of the American Society for Microbiology. p. 257. Moorman, T. B., J. M. Becerril, J. Lydon, and S. O. Duke. 1992. Production of hydroxybenzoic acids by Bradyrhizobium japonicum strains after treatment with glyphosate. Journal of Agricultural and Food Chemistry 40: 289-93. Glyphosate [N-(phosphonomethyl)glycine] effects on the growth and metabolism of the nitrogen-fixing soybean (Glycine max) symbiont Bradyrhizobium japonicum were investigated. In higher plants glyphosate inhibits the shikimic acid pathway, causing accumulations of shikimic acid and certain benzoic acids, such as protocatechuic (PCA) gallic acid. We sought to determine whether glyphosate also causes the accumulation of shikimic acid-derived benzoic acids in B. japonicum. Three strains (USDA) 110, 123, and 138) were grown in defined media lacking aromatic amino acids. Glyphosate at concentrations of 0.5 and 1 mM was inhibitory but not lethal. Cell death occurred in response to 5 mM glyphosate, and rapid death was observed at 10 mM. Accumulations of 1 mM PCA were observed in the culture media of strains 123 and 138. Only trace amounts of PCA accumulated in untreated controls or in strain 110 cultures. All strains were able to metabolize PCA in the absence of glyphosate. The high levels of PCA produced by B. japonicum strains may alter the symbiotic interactions between this bacterium and glyphosate-resistant varieties of soybean.

80.

81.

82.

114

Microflora and Fungi 83. Muller, M. M., C. Rosenberg, H. Siltanen, and T. Wartiovaara. 1981. Fate of glyphosate and its influence on nitrogen cycling in two Finnish agriculture soils. Bulletin of Environmental Contamination and Toxicology 27: 724-30. See Plant and Soil Residues Section. Olson, B. M., and C. W. Lindwall. 1991. Soil microbial activity under chemical fallow conditions effects of 2,4-D and glyphosate. Soil Biology and Biochemistry 23: 1071-76. Field and laboratory studies were made to examine the effects of 2,4-D and glyphosate on soil microbial activity under zero-tillage chemical fallow conditions. Glyphosate and 2,4-D applied three consecutive times at field rates and 10 times the field rates had no effect on microbial biomass-C, C mineralization or nitrification in the field. Laboratory studies showed that 2 and 100 times the field rate of 2,4-D reduced nitrification by 11 and 79%, respectively. In laboratory studies, 2,4-D incorporated into the soil reduced nitrification, but surface-applied 2,4-D did not. Glyphosate and 2,4-D used as recommended should not affect microbial activity under dryland zero-tillage chemical fallow conditions. Paula, Jr., T. J. D., and L. Zambolim. 1994. Effect of fungicides and herbicides on the establishment of the vesicular-arbuscular mycorrhizal fungus Glomus etunicatum on Eucalyptus grandis. Fitopatologia Brasileira 19, no. 2 : 173-77 . The effect of fungicides and herbicides was studied on the establishment of the vesiculararbuscular mycorrhizal fungus Glomus etunicatum on Eucalyptus grandis and on the growth of the seedlings in greenhouse. Benomyl, methyl thiophanate, PCNB and captan did not show any phytotoxic effect on the seedlings. Benomyl and methyl thiophanate reduced the colonization of E. grandis by G. etunicatum only at 250 ppm, but triadimefon, PCNB, mancozeb and glyphosate reduced it at 60, 55, 49 and 51%, respectively. Benomyl, triadimefon, methyl thiophanate, carboxin, iprodione, PCNB, captan, mancozeb, metalaxyl-mancozeb, glyphosate and oxyfluorphen were depressive to sporulation of G. etunicatum. (Brazil) . Paula, Jr., T. J. D., and L. Zambolim. 1994. Effect of fungicides and herbicides on the establishment of the ectomycorrhizal fungi Pisolithus tinctorius on Eucalyptus grandis . Summa Phytopathologica 20, no. 1: 30-34 . The goal of this study was to evaluate fungicides and herbicides on the in vitro mycelial growth of the ectomycorrhizal fungi P. tinctorius and on the establishment of the fungi on E. grandis in greenhouse. In solid and liquid MNM media, the fungicides triadimefon, mancozeb and metalaxylmancozeb inhibited completely the growth of P. tinctorius in vitro, in concentrations higher than 10 ppm. Carbaxin was depressive also in 1 ppm. Benomyl, methyl thiophanate, PCNB and captan did not show any phytotoxic effect on the seedlings. Benomyl did not affect the colonization of E. grandis by P. tinctorius. At concentration of 10 ppm, carboxin, iprodione, mancozeb, metalaxyl-mancozeb, glyphosate and oxyfluorphen reduced the colonization of the roots by P. tinctorius at 60, 55, 52, 48 and 44%, respectively. (Brazil) . Penaloza-Vazquez, A., G. L. Mena, L. Herrera-Estrella, and A. M. Bailey. 1995. Cloning and sequencing of the genes involved in glyphosate utilization by Pseudomonas pseudomallei. Applied and Environmental Microbiology 61, no. 2 : 538-43 . Thirty-four strains of Pseudomonas pseudomallei isolated from soil were selected for their ability to degrade the phosphonate herbicide glyphosate. All strains tested were able to grow on glyphosate as the only phosphorus source without the addition of aromatic amino acids. One of these strains, P. pseudomallei 22, showed 50% glyphosate degradation in 40 h in glyphosate medium. From a genomic library of this strain constructed in pUC19, we have isolated a plasmid carrying a 3.0-kb DNA fragment which confers to E. coli the ability to use glyphosate as a phosphorus source. This 3.0kb DNA fragment from P. pseudomallei contained two open reading frames (glpA and glpB) which are involved in glyphosate tolerance and in the modification of glyphosate to a substrate of the Escherichia coli carbon-phosphorus lyase. glpA exhibited significant homology with the E. coli hygromycin phosphotransferase gene. It was also found that the hygromycin phosphotransferase genes from both P. pseudomallei and E. coli confer tolerance to glyphosate. (Mexico) .

84.

85.

86.

87.

115

Microflora and Fungi 88. Petite, V., R. Cabridenc, R. P. J. Swannell, and R. S. Sokhi. 1995. Review of strategies for modelling the environmental fate of pesticides discharged into riverine systems. Environment International 21, no. 2: 167-76. See Water Quality Section. Petrie, G. A. 1995. Effects of chemicals on ascospore production by Leptosphaeria maculans on blackleg-infected canola stubble in Saskatchewan. Canadian Plant Disease Survey 75, no. 1 : 45-50. See Biodiversity and Habitat Restoration Section. Pipke, R., A. Schulz, and N. Amrhein. 1987. Uptake of glyphosate by an Arthrobacter sp. Applied Environmental Microbiology 53: 974-78. The uptake of glyphosate (N-[phosphonomethyl]glycine) by an Arthrobacter sp. which can utilize this herbicide as its sole source of phosphorus was investigated. Orthophosphate suppressed the expression of the uptake system for glyphosate and also compared with glyphosate for uptake. The Km for glyphosate uptake was 125 M, and the K, for orthophosphate was 24 M. Organophosphonates as well as organophosphates inhibited glyphosate uptake, but only organophosphates and orthophosphate suppressed the uptake system. Glyphosate uptake was energy dependent, had a pH optimum of 6 to 7, and was differentially affected by divalent cations. (Germany). Portier, R., K. Fujisaki, D. Friday, L. Hallas, and W. Adams. 1991. Continuous biotreatment of centrifuge spent wash effluents using immobilized microbial populations. 91st General Meeting of the American Society for Microbiology. p. 278. Powell, C. L., and D. J. Bagyaraj. 1985. Effects of some herbicides and fungicides on the in-vitro growth of the endomycorrhizal fungus Pezizella ericae. New Zealand Journal of Agricultural Research 27: 581-86. The endomycorrhizal fungus, P. ericae (Read), was grown in modified Norkran's broth amended with several rates of 2 herbicides (hexazinone and glyphosate) and 3 fungicides (etridiazole, vinclozolin and benomyl). At equivalent rates to those used in the field, there was no significant effect of hexazinone, glyphosate or etridiazole on fungal dry matter (DM) yields after 30 days incubation at 20 degrees Celcius. However, vinclozolin and benomyl reduced fungal DM production by 48% and 82%, respectively, although it is likely that P. ericae would be able to tolerate the low rates at which these fungicides would be present in plant roots under field conditions. (New Zealand). Powell, H. A., N. W. Kerry, and P. Powell. 1991. The mechanism of glyphosate resistance in the cyanobacterium Anabaena variabilis. British Phycological Journal 26: 94. (United Kingdom). Preston, C. M., and J. A. Trofymow. 1989. Effects of glyphosate (Roundup) on biological activity of two forest soils. Forest Resource Development Agreement. Forestry Canada, B.C. Ministry of Forests. FRDA Report 063, pp. 122-140. The effects of glyphosate (ROUNDUP) on some biological processes were examined in laboratory and field trials. In the laboratory study, forest floor material and mineral soil (0-5 cm depth) from a Douglas-fir site at Shawnigan Lake (Vancouver Island) were incubated with ROUNDUP (10 and 15 50 g a.i./g), with or without N -labelled urea (at 200 g/g N). No significant effects of ROUNDUP at either level of application were found on carbon dioxide evolution, urea hydrolysis, nitrification, or immobilization of ammonium, for either forest floor or mineral soil layers. In field studies, soil fauna and microflora populations were monitored in surface organic layers prior to and following ROUNDUP application on alder-covered sites at Carnation Creek (Vancouver Island). In a 6-month study, ROUNDUP had no significant longterm effects on populations of soil fauna or microflora. In an intensive 1-month field sampling trial, soil microflora populations in ROUNDUP-treated plots fluctuated and then returned to control levels.

89.

90.

91.

92.

93.

94.

116

Microflora and Fungi 95. 96. 97. Quilty, S. P., and M. J. Geoghegan. 1975. Effects of glyphosate on fungi. Proceedings of the Society for General Microbiology , II 87. . 1976. Effects of 'Roundup' on microbial populations in cultivated peat. Proceedings of the Society for General Microbiology , III 128. Quinn, J. P., J. M. M. Pedden, and R. E. Dick. 1988. Glyphosate tolerance and utilization by the microflora of soils treated with the herbicide. Applied Microbiology and Biotechnology 29: 511-16. The effect of recurrent applications of the herbicide glyphosate on a garden soil was investigated. Compared to an adjacent untreated soil the microbial population showed reduced sensitivity to glyphosate when grown in mineral salts medium. In both populations inhibition could be partially reversed by addition to the medium of the end products of the aromatic amino acid biosynthetic pathway, but the effect was more pronounced in the population from the treated site. However, all isolates from both soils were capable of growth in unsupplemented medium in the presence of as much as 10 mM glyphosate. No evidence for glyphosate metabolism was obtained from enrichment experiments carried out using inocular from the untreated soil; at the treated site organisms capable of using glyphosate as sole C or N source could not be isolated but a variety of gram-negative bacteria able to use its phosphonate moiety were obtained. Many of these organisms were identified as Pseudomonas spp. (N. Ireland). Quinn, J. P., J. M. M. Peden, and R. E. Dick. 1989. Carbon-phosphorus bond cleavage by grampositive and gram-negative soil bacteria. Applied Microbiology and Biotechnology 31: 283-87. Five soil bacterial isolates, originally selected for their ability to utilize the herbicide glyphosate as sole phosphorus source, were characterized with respect to their ability to use a range of other structurally-diverse phosphonates. Most showed broad substrate specificity and strains of Pseudomonas and of Bacillus megaterium were capable of degrading 14 of the other 15 phosphonates investigated. However no isolate was able to utilize isopropyl phosphonate, nor the phosphinate herbicide Phosphinothricin. Growth rates on most phosphonates were significantly lower than those sustained by inorganic phosphate and evidence was obtained for preferential utilization of the latter. In addition, the length of lag phase preceding growth on phosphonates varied widely. These characteristics are believed to reflect the diversity of routes by which such molecules enter bacterial cells and are metabolized. (N. Ireland). Rahe, J. E. Johal G. S., and A. Levesque. 1989. Synergistic role of soil fungi in the herbicidal activity of glyphosate. American Chemical Society , 197. Rahe, J. E., C. A. Levesque, and H. Speier. 1990. Evidence for specificities in the synergistic interaction of root-colonizing fungi and glyphosate. Canadian Journal of Plant Pathology 12: 339. Reynolds, P. E., J. A. Simpson, R. A. Lautenschlager, F. W. Bell, A. M. Gordon, D. A. Buckley, and D. A. Gresch. 1997. Alternative conifer release treatments affect below- and near-ground microclimate. Forestry Chronicle 73: (In press). Li-Cor weather stations and thermistor/resistance soil cells were used during 1994 to monitor microclimate in young spruce plantations during the first growing season after the following replicated alternative conifer release treatments (brush saw, Silvana Selective, Release [a.i., triclopyr] herbicide, Vision [a.i., glyphosate] herbicide), and control (no treatment) were applied. Treatments were conducted in mid-August (herbicides) and late October and early November (cutting) 1993. In 1994, temperature, photosynthetically-active radiation (PAR), and relative humidity (RH) were monitored near (0.25 m) and above (2 m) the forest floor. Fiberglass thermistor/resistance soils cells were installed 15 and 30 cm deep, and soil moisture and temperature were read bimonthly. In relation to controls, PAR near and above the forest floor increased on all conifer release treatments. By July, PAR near the forest floor declined on both the cut and herbicide treatments. That decline occurred in early July for the brush saw treatment, but in late July for the Vision treatment. PAR at 2 m was similar among conifer release alternatives and significantly greater than for controls throughout the growing season. Increased solar radiation resulted in significant soil warming following the conifer 117

98.

99. 100.

101.

Microflora and Fungi release treatments. During the growing season, duff (5 cm) and mineral (15 cm) soil temperatures were highest for the Vision and Release treatments, and lower on the brush saw and control treatments. November soil temperatures were slightly cooler in released than control plots. Frequent rains resulted in relatively high RH and soil moisture readings during the 1994 growing season. Relative humidity near the forest floor was lowest for the Vision, intermediate for the brush saw, and highest for the control treatments. During the growing season after treatments, soil moisture levels were higher on treated than control plots. 102. Rhodes, A. N., and C. W. Hendricks. 1990. A continuous-flow method for measuring effects of chemicals on soil nitrification. Toxicity Assessment 5: 77-90. Because the productivity of terrestrial ecosystems is directly related to microbial nutrient cycling, understanding the effects of chemical contaminants on soil microbial processes is important. This study examined the effects of two model chemicals-Roundup (glyphosate) and N-Serve (nitrapyrin)-on nitrifying organisms in static, perfusion, and continuous-flow culture systems. Experimental concentrations were approximately 1, 10, and 100 times the spot application rate. Both N-Serve and Roundup were shown to inhibit nitrification in the treated soils. Roundup significantly reduced nitrification at 6.8 and 68 mg/g dry soil. N-Serve (nitrapyrin) completely inhibited nitrification at levels greater than 42 m/g dry soil in all cultural methods. In comparative studies with static batch and perfusion culture techniques, the continuous-flow system proved to be both reliable and useful in the culture of nitrifying bacteria. This method provides an alternative to traditional culture techniques in measuring chemical effects on microbial geochemical cycles and provides a new method for use in toxicity testing. Robertson, B., and M. Alexander. 1994. Growth-linked and cometabolic biodegradation: Possible reason for occurrence or absence of accelerated pesticide biodegradation. Pesticide Science 41, no. 4 : 311-18 . A study was conducted to relate the occurrence of accelerated pesticide biodegradation to the susceptibility of the pesticides to growth-linked degradation or cometabolism. The mineralization of 2,4-D was initially slow but then became rapid, and a second application was mineralized with no acclimation phase and more rapidly than the first. The numbers of 2,4-D-degrading micro-organisms increased markedly following its first application and then declined, but the population size increased after a second addition. Glyphosate was rapidly and extensively mineralized following the first and second applications to soil, and the abundance of organisms able to degrade it rose after the first addition and remained high before and following the second application. Propham (IPC) mineralization was detected only 15 days after its application but the degradation was rapid thereafter, and the second addition was rapidly and extensively mineralized with no acclimation phase. The population of propham-degrading micro-organisms was initially small, but increased markedly 10 days after the initial herbicide addition and was still large at the time of the second application. The rate of carbofuran biodegradation in the test soil was the same following the first and second applications, and the abundance of carbofuran-metabolizing microorganisms did not change appreciably as a result of soil treatment with the insecticide. Simazine mineralization was slow, although the rate was higher following the second addition; however, the number of simazine-degrading organisms did not increase appreciably. From 10 to 12% of the 14C from radiolabeled 2,4-D, propham, glyphosate or glucose was usually incorporated into the microbial biomass of soil but 0.82% or less of the 14C from simazine or ring- or carbonyl-labeled carbofuran was converted to biomass. It is suggested that pesticides that support microbial growth may be subject to accelerated biodegradation if the population remains large until the pesticide is applied again. Pesticides that do not support growth may not be subject to accelerated biodegradation. Roslycky, E. B. 1982. Glyphosate and the response of the soil microbiota. Soil Biology and Biochemistry 14: 87-92. Limited information on the effect of glyphosate (N-phosphonomethylglycine) on soil microorganisms justified an inquiry into the response of soil actinomycetes, bacteria and fungi in terms of their respiration, and sensitivity of isolates. Low concentrations of glyphosate had little effect on total populations of these organisms during the 214-day experiment, while high concentrations initially increased actinomycete and bacterial numbers by 2 and 1 logs, respectively. The stimulation was

103.

104.

118

Microflora and Fungi followed by a decline and fluctuation showing a gradual increase in numbers. The respiration rates of the soil microbiota in soil suspensions, showed some irregular stimulation and retardation with up to 10 m.g glyphosate/ml. In contrast high doses suppressed 02 uptake by the microbiota. Fungi were the least affected. Pronounced inhibition of actinomycete and bacterial respiration was in agreement with the results from isolate replication. The results indicated both stimulation and inhibition of 02 uptake by some organisms within these groups. In contrast to some reports of limited, shortterm inquiries these results showed considerable effects of glyphosate on soil microorganisms. 105. Rueppel, M. L., B. B. Brightwell, M. Schaefer, and J. T. Marvel. 1977. Metabolism and degradation of glyphosate in soil and water. Journal of Agricultural and Food Chemistry 25: 517-28. See Plant and Soil Residues Section. Sahid, I., M. A. Salleh, and O. Omar. 1993. Effects of herbicides on cellulose decomposition in soil. Plant Protection Quarterly 8, no. 2: 38-39. (Malaysia). Santos, A., and M. Flores. 1995. Effects of glyphosate on nitrogen fixation of free-living heterotrophic bacteria. Letters in Applied Microbiology 20 , no. 6 : 349-52 . The effect of the herbicide glyphosate (N-(phosphonomethyl) glycine) on the growth, respiration and nitrogen fixation of Azotobacter chroococcum and A. vinelandii was studied. A. vinelandii was more sensitive to glyphosate toxicity than A. chroococcum. Recommended dosages of glyphosate did not affect growth rates. More than 4 kg/ha is needed to find some inhibitory effect. Specific respiration rates were 19.17 mmol 02/h/g dry weight for A. chroococcum and 12.09 mmol/h/g for A. vinelandii. When 20 kg/ha was used with A. vinelandii, respiration rates were inhibited 60%, the similar percentage inhibition A. chroococcum showed at 28 kg/ha. Nitrogen fixation dropped drastically 80% with 20 kg/ha in A. vinelandii and 98% with 28 kg/ha in A. chroococcum. Cell size as determined by electron microscopy decreased in the presence of glyphosate, probably because glyphosate induces amino acid depletion and reduces or stops protein synthesis. (Spain) . Schuster, E., and D. Schroder. 1990. Side-effects of sequentially- and simultaneously-applied pesticides on non-target soil microorganisms: laboratory experiments. Soil Biology and Biochemistry 22: 375-83. In laboratory experiments the herbicides glyphosate (Roundup) and dichlorprop (U-46 DP) both at 4 kg/ha, the fungicides prochloraz + carbendazim (Sportak Alpha) at 1.5 kg, captafol + triadimefon (Bayleton DF) at 4 kg, an insecticide pirimicarb (Pirimor) at 0.3 kg and the growth regulator chlormequat (Cyclocel) at 0.5 kg induced only minor effects in microbial activity in a parabrown soil. Bayleton, Primidor and Roundup showed interactions that were not consistent with time. The additive effect of Bayleton DF and Roundup was more inhibitory than the combination of all 3 substances. At high dosages, side-effects were completely dominated by the strong and irreversible inhibitions exerted by Bayleton. Sequential applications of pesticides often resulted in stronger side-effects as compared to simultaneous applications. The results are discussed with respect to an assessment of the possible hazards due to multiple pesticide applications in modern agriculture. (Germany). . 1990. Side-effects of sequentially-applied pesticides on non-target soil microorganisms: field experiments. Soil Biology and Biochemistry 22: 367-73. In trials in 1985-86, the effect of a plant protection system consisting of 7 pesticide treatments was measured on the microbial flora of a Gleyic Luvisol soil. Successive applications of 3 kg dinosebacetate (Aretit)/ha, 2.5 kg isoproturon (Arelon), 4 kg dichlorprop (U-46 DP), 0.5 kg chlormequat (Cyclocel), 1.5 kg prochloraz + carbendazim (Sportak Alpha), 4 kg captafol + triadimefon (Bayleton DF), 0.3 kg pirimicarb (Pirimor) and 4 kg glyphosate (Roundup) caused only slight and short-lived sideeffects on microbial activity which usually disappeared before the next treatment was applied. The impact of the pesticides on soil microflora appeared to be significantly affected by climatic conditions. (Germany).

106.

107.

108.

109.

119

Microflora and Fungi 110. Selvapandiyan, A., and R. K. Bhatnagar. 1994. Cloning of genes encoding for C-P lyase from Pseudomonas isolates PG2982 and GLC11: Identification of a cryptic allele on the chromosome of P. aeruginosa. Current Microbiology 29, no. 5 : 255-61. Two isolates of Pseudomonas sp., GLC11 and PG2982, can use glyphosate as a sole source of phosphorus. This ability is indicative of enzymatic cleavage of a carbon-phosphorus bond, and the enzyme has been named C-P lyase. We have cloned, in Escherichia coli, gene/s coding for C-P lyase on a broad host range cosmid pLA2917. Restriction fragment arrangement of cloned fragments of PG2982 and GLC11 has been established. Analysis by Southern hybridization between two clones revealed a strong homology between three PstI fragments of pPG-CP-14 (derived from PG2982) and pGC-CP-4 (derived from GLC11). With the construct pGC-CP-4 as a probe, the presence of a cryptic allele for C-P lyase has been demonstrated on the chromosome of the parent isolate, Pseudomonas aeruginosa PAO1. It is suggested that genetic rearrangement such as frame shift or a point mutation activated the cryptic C-P lyase gene. Metabolism of glyphosate by E. coli carrying pPG-CP-14 or pGCCP-4 has been demonstrated by radiometric experiments. (India) . . 1994. Isolation of a glyphosate-metabolizing Pseudomonas: Detection, partial purification and localization of carbon-phosphorus lyase. Applied Microbiology and Biotechnology 40, no. 6 : 876-82 . A Pseudomonas isolate (GLC11) capable of growth in the presence of up to 125 mM glyphosate (N-phosphonomethyl glycine (PMG)) has been isolated. Unlike the previously isolated Pseudomonas PG2982 and other bacterial strains, isolate GLC11 grows equally well in commercial formulation and analytical grade PMG. Utilization of PMG as a phosphorus source is repressed by inorganic phosphate (Pi) in both isolates. Enzymatic activity responsible for carbon-phosphorus bond cleavage (C-P lyase) was detected in cell-free extracts of both isolates and was partially purified. Resolution on DE-52 anion exchange chromatography yielded a single peak of C-P lyase activity. The molecular mass of C-P lyase as analysed by gel permeation chromatography is approximately 200 kDa. The enzyme activity was localized in the periplasmic space of bacteria. The specific activity of C-P lyase was different for different phosphonates when used as substrates. (India) . Sharma, U., E. A. Adee, and W. F. Pfender. 1989. Effect of glyphosate herbicide on pseudothecia formation by Pyrenophora tritici repentis in infested wheat straw. Plant Disease 73: 647-50. Under moist conditions in the greenhouse, formulated herbicides containing bromoxynil, dicamba, glyphosate, 2,4-D or paraquat, applied at labelled rates, significantly reduced ascocarp production by P. tritici repentis. No ascocarps were produced in straw treated with the glyphosatecontaining herbicide Roundup. In further experiments with this herbicide, greenhouse-grown straw infested with P. tritici repentis was treated either before incubation under conditions conducive to ascocarp development or at one of several times after incubation had started. The herbicide completely inhibited ascocarp development if applied before the conducive environment was imposed, but this inhibition diminished as the time delay increased between inoculation and treatment. Experiments conducted on autoclaved, inoculated straw indicated that the herbicide does not greatly reduce the mycelial growth rate of P. tritici repentis. It was not determined whether the effect of the formulated material is due to glyphosate or to the non-herbicidal components of the material. Sidhu, S. S., and P. Chakravarty. 1990. Effect of selected forestry herbicides on ectomycorrhizal development and seedling growth of lodgepole pine and white spruce under controlled and field environment. European Journal of Forest Pathology 20: 77-94. In tests under aseptic conditions, the herbicides hexazinone (PrononeTM5G and Velpar L.R.), glyphosate, and triclopyr reduced seedling growth and mycorrhizal development of Pinus contorta var. latifolia and Picea glauca. Triclopyr was most toxic of the four herbicide formulations. Under greenhouse condition, only 2 formulations of hexazinone (PrononeTM5G and Velpar L.) were tested. At high concentrations (2 and 4 kg/ha) it reduced growth and mycorrhizal infections significantly but showed recovery with time. No adverse effects were observed at low concentration (1 kg/ha). In general, seedlings inoculated with mycorrhizal fungus (Suillus tomentosus), were more sensitive to herbicide than the ones without mycorrhizal inoculation. Under field conditions overall effects of herbicide applicated were less intense. Only 4 kg/ha rates of hexazinone resulted in reductions in seedling growth and mycorrhizal infections. The nonmycorrhizal seedings planted in the field 120

111.

112.

113.

Microflora and Fungi developed mycorrhizae in over 40% of short roots within 2 months after planting in all herbicide treatments. 114. Soulas, G. 1992. Biological availability of pesticides in soil 2,4-D and glyphosate as test cases. Proceedings of the International Symposium on Environmental Aspects of Pesticide Microbiology J.P.E. Anderson et al. (editors). pp. 219-224. (Sweden). Spurgeon, D. W., and A. J. Mueller. 1991. Temporal effects of threecornered alfalfa hopper (Homoptera: Membracidae) girdling on translocation in soybean. Journal of Economic Entomology 84: 1203-7. The temporal effects of petiole girdles made by 1-, 7- and 14-day-old Spissistilus festinus on translocation of soyabean were investigated using the systemic herbicide glyphosate as a tracer. Leaves of S. festinus-girdled, stem-girdled, and ungirdled petioles were treated with glyphosate. Diffusive resistance, leaf temperatures and stem growth rates of treated plants were compared with those of untreated plants to detect translocation of the herbicide. S. festinus girdles initially blocked translocation. Blockage was followed by recovery of phloem function in 7 days. Resumption of translocation was not due to apoplastic transport. Stratton, G. W., and K. E. Stewart. 1991. Effects of the herbicide glyphosate on nitrogen cycling in an acid forest soil. Water, Air and Soil Pollution 60: 231-48. The herbicide glyphosate was sprayed aerially on a section of conifer forest in Atlantic Canada that had been previously clearcut and reforested. Glyphosate was then tested for effects on ammonification, nitrification, and denitrification for a period of 8 months by comparing microbial activity in treated and untreated zones of the clay loam forest soil and the overlying decomposing litter, both with a pH of 3.8. With ammonification, there was generally a stimulation of activity in both the forest litter (FL) and forest soil (FS) that had been exposed to glyphosate during spraying. Nitrification rates in FL and FS were very low and glyphosate had no appreciable stimulatory or inhibitory effect on nitrification. Although glyphosate stimulated denitrification in a few instances, it generally had no significant effect on denitrification activity in FL and FS exposed during spraying. With all processes, microbial activity in FL was significantly greater than that in FS. Laboratory bioassays were also performed with FL and FS, as well as two silt loam (pH 5.8 and 6.4) and one sandy loam (pH 6.8) agricultural soils, using glyphosate concentrations up to 200 times higher than field application rates. With ammonification and denitrification, glyphosate generally stimulated activity at all levels tested and in all soil used. Glyphosate stimulated ammonification by 50% at concentrations ranging from 140 to 550 mg/g for the soils and >4000 mg/g for FL. With denitrification, the corresponding herbicide levels were approximately 2250 mg/g for FS, >10,000 for FL, and 450 for an agricultural soil. With nitrification, it was estimated that glyphosate concentrations greater than 1000 to 2000 mg/g would be required to cause a 50% inhibition of activity. The careful use of glyphosate in forestry should have no toxic effects on N cycling in soils. . 1992. Glyphosate effects on microbial biomass in a coniferous forest soil. Environmental Toxicology and Water Quality 7: 223-36. The herbicide glyphosate (Roundup) was applied aerially to a conifer forest that had previously been clear-cut and reforested. Glyphosate was tested for effects on microbial biomass, numbers of selected microorganisms, and soil respiration over a period of 8 months by comparing treated and untreated zones of the clay loam forest soil (FS) and the overlying litter (FL), both pH 3.8. With microbial biomass, glyphosate generally had a stimulatory effect in FL, but usually no significant effect in FS. Glyphosate had no significant effect on numbers of bacteria, fungi, and actinomycetes in either FL or FS. The herbicide generally stimulated respiration in both FL and FS. In laboratory bioassays using FL, FS, and three agricultural soils, glyphosate had no significant effect on respiration when used at concentrations up to 100 times higher than recommended field application rates. Respiration rates in FL, as measured with Warburg respirometry, were unaffected by glyphosate. Respiration rates in FS were stimulated by glyphosate concentrations 10 and 100 times higher than recommended field rates. Glyphosate should have no deleterious effects on microbial biomass and respiration in forest soils when used under recommended conditions.

115.

116.

117.

121

Microflora and Fungi 118. Su, C. Y. 1988. The effect of certain pesticides on Beauveria bassiana. Chinese Journal of Entomology 8: 157-60. The effects of Tokuthion (prothiofos), Marshal (carbosulfon), mevinphos, paraquat, glyphosate, Stomp (pendimethalin), Previcura (propamocarb), Terrazole (etridiazole) and Sportak (prochloraz) at dilutions of 500, 1000 and 1500 times on the mycelial growth of Beauveria bassiana were studied in the laboratory. Sportak 25% EC was the most toxic, followed by Previcura. Sportak inhibited fungal growth at all dilutions tested. Paraquat was the least toxic of the pesticides tested. (China). Tooby, T. E. 1985. Fate and biological consequences of glyphosate in the aquatic environment. In: The Herbicide Glyphosate. E. Grossbard, and D. Atkinson, London: Butterworths. pp. 20617. See Water Quality Section. Torstensson, L. 1978. Glyphosate degradation and effects on soil microorganisms. Weeds and Weed Control 19: 78. . 1985. Behaviour of glyphosate in soils and its degradation. In: The Herbicide Glyphosate. Edited by E. Grossbard, and D. Atkinson, London: Butterworths. pp. 137-49. See Plant and Soil Residues Section. Torstensson, N. T. L., and A. Aamisepp. 1977. Detoxification of glyphosate in soil. Weed Research 17: 209-12. Detoxification of glyphosate [N-(phosphonomethyl)-glycine) in nonsterile and autoclaved soils was followed by bioassay with wheat. Comparisons were made with detoxification of MCPA [2-methyl4-chlorophenoxyacetic acid] under similar conditions followed by bioassay with spring rape. The well known pattern for microbial metabolism of MCPA with a lag phase preceding the rapid degradation was shown. The initial rapid inactivation of glyphosate is by adsorption, but the results also clearly indicate that the further disappearance of activity depends mainly on microbial degradation. Glyphosate does not seem to sustain microbial growth, which indicates that it is degraded by cometabolism. In autoclaved soil the possibility of a slight chemical degradation or an adsorption that becomes stronger with time could not be excluded. (Sweden). Toubia-Rahme, H., D. E. Ali-Haimoud, G. Barrault, and L. Albertini. 1995. Inhibition of Drechslera teres sclerotioid formation in barley straw by application of glyphosate or paraquat. Plant Diseases 79, no. 6 : 595-98 . Field-grown barley straw was inoculated with Drechlsera teres f. sp. teres or D. t. maculata, treated with two herbicides used in no-tillage barley production, and then incubated in controlled conditions to induce sclerotioid structure morphogenesis (resting form). Formulated herbicides containing glyphosate or paraquat were applied at three different concentrations. Applied at the recommended field rate, these herbicides significantly reduced sclerotioid structure production by D. teres. In addition, their morphology and myceliogenesis were modified in the presence of both herbicides at the recommended field rate. Glyphosate was more inhibitory than paraquat; no sclerotioid structures were produced in straw when glyphosate was applied before colonization by D. teres. The effect of the herbicides at 10 super(-8) and 10 super(-4) M varied depending on the herbicide concentration of the active ingredient, method of application, and forma specialis of the pathogen. (France) .

119.

120. 121.

122.

123.

122

Microflora and Fungi 124. Tropea, M., G. Fisichella, and A. Longo. 1979. Influence of glyphosate on CO2 evolution in some typical soils of eastern Sicily. Tecnica Agricola 31, no. 1-3: 11 . (Italy). Tu, C. M. 1994. Effects of herbicides and fumigants on microbial activities in soil. Bulletin of Environmental Contamination and Toxicology 53, no. 1 : 12-17 . Uotila, M., T. Komives, P. Ott, and Z. Klement. 1994. Effects of glyphosate on compatible and incompatible plant-bacterium interactions. Acta Phytopathologica Et Entomologica Hungarica 29, no. 3-4 : 353-59 . Effects of the herbicide glyphosate on the infection by Pseudomonas bacteria of susceptible and resistant bean varieties were investigated. Glyphosate, up to 250 millimolar concentrations had no effect on the growth of Pseudomonas syringae pv. phaseolicola bacteria in vitro. Subphytotoxic levels of glyphosate did not influence the compatible interaction between the bacteria and the susceptible trifolia of Red Kidney bean plants. Following treatment with glyphosate, the type of the incompatible interaction between the bacteria and the trifolia and primary leaves of the resistant bean variety O2, and primary leaves of older Red Kidney bean plants was shifted towards the compatible one. (Finland) . Vannini, C., M. C. Napoli, N. Miclaus, E. Casalone, and E. Gallori. 1990. Influence of different pesticides on Azospirillum brasilense and Azotobacter chroococcum and microbial processes related to the mechanism of detoxification. Agrokemia Es Talajtan 39: 503-8. The effect of 5 herbicides (atrazine, EPTC, glyphosate, metolachlor and 2,4-D), 12 fungicides and 2 insecticides (butylate and pyrethrines) on the growth, nitrogenase activity and glutathione content of Azospirillum brasilense str. ATCC 29710 and Azotobacter chroococcum str. AzWT are reported. The fungicides captan and ziram were the most toxic of the compounds assayed. Additions of cysteine and glutathione to the medium nullified the toxicity of the 2 fungicides. 1,8-Naphthalic anhydride antidote, however, enhanced the fungicides' toxicity when added at sublethal concn to A. brasilense and A. chroococcum cultures. (Italy). Vasilev, K. 1982. Glyphosate effect on test microorganisms under laboratory conditions. Khigiena I Zuraveopazvane 25: 346-51. The effect of the herbicide glyphosate was studied on some representative species of water saprophytic microflora and of some sanitary-indicative and pathogenic intestinal microorganisms. Strains from the following species were used: Pseudomonas aeruginosa, Escherichia coli and Salmonella typhimurium. The experiments were carried out under laboratory conditions with and without aeration, with 3 concentrations of glyphosate (10, 100 and 300 mg/dm3). The quantitative changes in the microorganisms were studied by inoculations on solid nutritive media. Glyphosate effect 3 on test microorganisms was concentrationdependent. At a concentration of 10 mg/dm , glyphosate 3 induced an increase in the quantity of the test microorganisms, at a concentration of 300 mg/dm , an inhibitory effect was established, manifested by a reduction of the number of microorganisms. All tested glyphosate concentrations, in combination with aeration, induced a sharp decrease in the quantity of the test microorganisms. (Bulgaria). Wallace, R. W., and R. R. Bellinder. 1993. Rust-infected quackgrass (Elytrigia repens) growth and rhizome bud death after glyphosate application. Weed Science 41 , no. 3 : 501-7 . Photosynthesis decreased 32% in quackgrass infected with rust. Photoassimilate transport in rust-infected quackgrass, assessed by autoradiography, was less than that of control (no rust) leaves 35 min after exposure to 14CO2. Greenhouse-grown quackgrass height, foliar and rhizome dry weights, rhizome length, and bud numbers decreased 11, 25, 28, 19, and 21% in rust-infected plants, respectively, compared to control plants. Foliar and rhizome biomass decreased 31 and 39%, respectively, in field-grown rust-infected plants. Rhizome bud death 48 h after treatment (HAT) in both control and rust-infected plants increased as glyphosate rate increased from 0.13 to 1 kg ae/ha; however, control plant bud death was significantly greater than that of rust-infected plants.

125. 126.

127.

128.

129.

123

Microflora and Fungi 130. Wardle, D. A., K. S. Nicholson, and A. Rahman. 1994. Influence of herbicide applications on the decomposition, microbial biomass, and microbial activity of pasture shoot and root litter. New Zealand Journal of Agricultural Research 37, no. 1 : 29-39 . Pure swards of each of four pasture species (Lolium perenne L., Trifolium repens L., Senecio jacobaea L., and Carduus nutans L.) were established in glasshouse conditions and subjected to one of three treatments: spraying with 2,4-D/picloram mix; spraying with glyphosate; or unsprayed. After the sprayed swards died, all above-ground and below-ground tissue was harvested, air-dried, and placed in nylon mesh litter-bags which were positioned in the field. Decomposition, microbial basal respiration, and substrate-induced respiration (proportionally related to the glucose-responsive microbial biomass) of this litter was then monitored over 338 days. Both herbicide treatments inhibited decomposition of T. repens and L. perenne shoot tissue and C. nutans root tissue, but stimulated that of C. nutans shoot tissue, indicating that herbicides may influence decomposition of different species in different ways; the possible reasons for this are discussed. However, the rapid decomposition of most of the tissues considered in this study suggest that herbicides are unlikely to exert substantial long-term effects on plant litter persistence. Microbial basal respiration and substrate-induced respiration of most of the litter types considered were initially very strongly enhanced by both herbicide treatments; however, this effect was highly transitory for all tissue types except one, and for some of the tissue types a strong inhibition of these microbial variables in the herbicide treatments followed. It therefore appears that microbial build-up on litter from herbicide-killed plants (and the subsequent decline) occurs earlier than that from unsprayed plants, probably because herbicide-induced plant damage increases the availability of readily utilisable microbial substrates. The retardation of leaf litter decomposition in herbicide treatments was often associated with reduced microbial activity and biomass, indicating strong linkages between soil-associated microflora and decomposition processes. This study also indicates that newly developed approaches for simultaneously assessing decomposition and the microbial biomass of leaf litter have considerable potential for investigating impacts of ecological factors on plant litter-microbial interactions. (New Zealand) . Wardle, D. A., and D. Parkinson. 1990. Effects of three herbicides on soil microbial biomass and activity. Plant Soil 122: 21-28. Three post-emergence herbicides (2,4-D, picloram and glyphosate) were applied to samples of an Alberta (Canada) agricultural soil at concentrations of 0, 2, 20, and 200 m.g/g. The effects of these chemicals on certain microbial variables was monitored over 27 days. All herbicides caused enhancement of basal respiration but only for 9 days following application, and only for concentrations of 200 m.g/g. Substrate-induced respiration was temporarily depressed by 200 m.g/g picloram and 2,4-D and briefly enhanced by 200 m.g/g glyphosate. It is concluded that because changes in microbial variables only occurred at herbicide concentrations of much higher than that which occurs following field application, the side-effects of these chemicals is probably of little ecological significance. . 1990. Influence of the herbicide glyphosate on soil microbial community structure. Plant Soil 122: 29-38. The side effects of glyphosate on the soil microflora were monitored by applying a range of glyphosate concentrations (0, 2, 20, and 200 m.g/g herbicide) to incubated soil samples, and following changes in various microbial groups over 27 days. Bacterial propagule numbers were temporarily enhanced by 20 m.g/g and 200 m.g/g glyphosate, while actinomycete and fungal propagule numbers were unaffected by glyphosate. The frequency of three fungal species on organic particles in soil was temporarily enhanced by 200 m.g/g glyphosate, while one was inhibited. One species was temporally enhanced on mineral particles. However, many of these fungi were inhibited by 200 m.g/g glyphosate in pure culture. There was little agreement between species responses to glyphosate in incubated soil samples and in pure culture. . 1991. Relative importance of the effect of 2,4-D, glyphosate and environmental variables on the soil microbial biomass. Plant Soil 134: 209-20. Two post-emergence herbicides (glyphosate and 2,4-D) were applied at field application levels to tilled field plots in a mixed cropping area in south-central Alberta. The effects of these chemicals on certain variables associated with microbial biomass and activity were monitored in these plots (as well 124

131.

132.

133.

Microflora and Fungi as corresponding control plots) for 45 days. Glyphosate did not influence any of the microbial variables tested. Addition of 2,4-D significantly influenced all microbial variables investigated but these effects were transient, being detectable only within the first 1-5 days of herbicide addition. The effects of 2,4-D addition on the microbial variables tested, even when significant, were typically small and probably of little ecological consequence especially when spatial and temporal variation in these variables is taken into account. 134. . 1992. The influence of the herbicide glyphosate on interspecific interactions between four soil fungal species. Mycological Research 96: 180-186. The influence of the herbicide glyphosate on two-way interactions between four fungal species (Mucor hiemalis, Mortierella alpina, Trichoderma harzianum, Arthrinium sphaerospermum) was investigated on agar using direct opposition, volatile inhibitor and non-volatile inhibitor tests, and on barley straw using tests for competitive saprotrophic ability. The results indicated that glyphosate was often capable of shifting the direction of two-species interactions, but there was little agreement between the different tests. Reanalysis of particle colonization data presented earlier demonstrated that glyphosate frequently eliminated negative associations between the same four species, possibly by acting as a resource. It is concluded that interactions between pairs of fungi can be influenced directly by glyphosate (especially at high concentrations) and that this may influence soil fungal community structure. . 1992. Influence of the herbicides 2,4-D and glyphosate on soil microbial biomass and activity: a field experiment. Soil Biology and Biochemistry 24: 185-86. Field trials were conducted at a site in S.-central Alberta to evaluate the effects on soil microbial basal respiration (BR) and substrate-induced respiration (SIR) of the following treatments: manual weeding every 3-6 d vs. no weeding, sowing with barley cv. Klondyke at 180 kg seed/ha vs. no crop, and treating with 4 kg/ha 2,4-D or 5 kg glyphosate vs. no herbicide. BR was unaffected by 2,4-D, except after 4 d in barley-sown non-weeded plots, and after 291 d in weeded plots with no crop. The effect of glyphosate on BR was not significant in any plot. 2,4-D did not exert any consistent effects on SIR. However, SIR values were significantly reduced (P = 0.01) by glyphosate after 8 d on plots with weeds but without barley plants. It was concluded that both herbicides were capable of reducing microbial activity or biomass within the 1st few d following application, but only in plots with weeds present. Weidhase, R., B. Albrecht, M. Stock, and R. A. Weidhase. 1990. Utilization of glyphosate by Pseudomonas sp. GS. Zentralbl Mikrobiol 145: 433-38. Pseudomonas sp. GS grown in the presence of 14C-labelled glyphosate metabolizes the herbicide to inorganic phosphate as well as in respect of utilization of the C-skeleton. Furthermore it was shown that the primary step of degradation is the breakdown of the C.sbd.P.sbd.bond, because sarcosine was found as a transient metabolite. (Germany). Wong, P. T. W., P. M. Dowling, L. A. Tesoriero, and H. I. Nicol. 1993. Influence of preseason weed management and in-crop treatments on two successive wheat crops: 2. Take-all severity and incidence of rhizoctonia root rot. Australian Journal of Experimental Agriculture 33, no. 2 : 173-77 . The effects of cultivation and herbicide use to control weeds in wheat on wheat growth, the severity of take-all, and the incidence of rhizoctonia root rot were studied for 2 seasons. Preseason treatments were no weed control, paraquat (0.20 kg a.i./ha), glyphosate (0.18 kg a.i./ha or 4 applications of 0.72 kg a.i./ha), and heavy grazing. In-crop treatments were cultivation plus trifluralin, direct drilling plus chlorsulfuron, and direct drilling alone. At the site, take-all was the main disease while rhizoctonia root rot was relatively minor. Glyphosate applied 4 times at 0.72 kg a.i./ha over the previous spring and summer led to greater wheat dry matter (DM) production, significantly (P lt 0.05) less severe take-all, and a lower incidence of rhizoctonia root rot in the first year than the other preseason treatments. Spraytopping with glyphosate (0.18 kg a.i./ha) or paraquat (0.20 kg a.i./ha) and heavy grazing reduced take-all severity but not the incidence of rhizoctonia root rot. Conventional cultivation resulted in more wheat DM, significantly less severe take-all, and a lower incidence of rhizoctonia root rot than direct drilling. Grain yields reflected the trends of the DM production despite 125

135.

136.

137.

Microflora and Fungi severe yield loss due to head frosting. Plots were split for cultivation and direct drilling in the second year. The highest wheat DM and grain yields were in the cultivated treatments but the effects of cultivation on take-all did not carry over from the first year. In both years, take-all was most severe in the control treatment and least severe in the treatment with the high rate of glyphosate (P lt 0.05). In the second wheat crop, however, take-all severity was similar in the 2 glyphosate, paraquat, and grazed treatments. The effect of a weed-free fallow obtained by use of a high rate of glyphosate was nullified in the second wheat crop because of a high carryover of volunteer wheat seedlings during the intervening wet summer. There was also a greater incidence of rhizoctonia root rot in the control than in the other treatments, and cultivation again reduced disease incidence compared with direct drilling. (Australia) . 138. Yasem de Romero, M. G. 1986. Effect of some agricultural chemicals on the entomopathogenic fungus Verticillium lecanii (Zimm.) viegas. Revista De Investigacion, Centro De Investigaciones Para La Regulacion De Poblaciones De Organismos Nocivos 4: 55-62. The effects of several pesticides currently used in citrus in Argentina on the growth and development of the entomogenous fungus Verticillium lecanii were evaluated in the laboratory. All 6 fungicides tested inhibited growth and conidia production. The tested insecticides also showed inhibitory action in the order chlorpyrifos dicofol=tetradifon chlorobenzilate. The herbicide glyphosate had no effect on conidia production and only reduced mycelial growth by 17%, while paraquat significantly reduced both. (Argentina). Zaranyika, M. F., and M. G. Nyandoro. 1993. Degradation of glyphosate in the aquatic environment: An enzymatic kinetic model that takes into account microbial degradation of both free and colloidal (or sediment) particle adsorbed glyphosate. Journal of Agricultural and Food Chemistry 41, no. 5: 838-42. See Water Quality Section.

139.

126

Plant and Soil Residues

Plant and Soil Residues


1. Ana'yeva, N. D., B. P. Strekozov, and O. K. Tyuryukanova. 1987. Change in the microbial biomass of soils caused by pesticides. Soviet Soil Science 18: 56-62. (Soviet Union). Aubin, A. J., and A. E. Smith. 1992. Extraction of Carbon-14-labelled glyphosate from Saskatchewan soils. Journal of Agriculture and Food Chemistry 40: 1163-65. The extraction of 14C from air-dried samples of three soils fortified 14 days previously with (14C)-glyphosate was compared using 10 different solvent systems and an extended-shaking procedure. 14C recoveries exceeding 73% were reproducibly achieved with 0.35 M H3PO4 (and 0.09 M with respect to CaCl2), 0.1 M NaOH, and 0.5 M NH4OH. With the phosphoric acid and ammonium hydroxide extractants highest recoveries from a loamy sand were recorded with a soil/solvent ratio of 1:5 or 1:10. With the aqueous sodium hydroxide, highest recoveries were achieved from all soils using a soil/solvent ratio of 1:2.5. Recoveries of 14C from all three soils, using 0.1 M NaOH, were not significantly different whether fortified 14, 28, or, in the case of clay, 56 days previously with (14C)glyphosate. Some loss of extractable radioactivity was noted in two of the soils after 56 days. Bowmer, K. H., P. M. D. Boulton, D. L. Short, and M. L. Higgins. 1986. Glyphosate-sediment interactions and phytotoxicity in turbid water. Pesticide Science 17: 79-88. In aquatic situations it was expected that adsorption by sediments and suspended solids would influence the movement of glyphosate away from the application zone and attenuate its phytotoxicity. However, the results of 2 experiments showed that only a minor proportion of glyphosate was adsorbed onto suspended solids even in turbid irrigation water. Phytotoxicity, as measured by the effect on the root growth of safflower, was not significantly reduced. Where glyphosate was intentionally injected into flowing water of contrasting quality to simulate incidental contamination of water during foliage treatment, adsorption by benthic sediments attenuated loads of glyphosate only slowly. The attenuation was 13-27% for each kilometre of travel downstream compared with 31% observed previously. However, when glyphosate was sprayed onto the sediment exposed after channel draining, 7% of the glyphosate applied was subsequently eluted. Consequently, draining before treatment should be an effective strategy for minimizing the contamination of irrigation water. (Australia). Carlisle, S. M., and J. T. Trevors. 1986. Effect of the herbicide glyphosate on nitrification denitrification and acetylene reduction in soil. Water, Air, and Soil Pollution 29: 189-204. The effect of glyphosate on soil respiration and H2 oxidation in an agricultural soil was investigated. The effects of the pure herbicide and commercial formulation, RoundupR (Monsanto Company), were compared in soil under both aerobic and anaerobic conditions. Both formulations stimulated 02 uptake as well as aerobic and anaerobic C02 evolution. Roundup caused more stimulation than glyphosate under aerobic incubation conditions; the formulations had an equal effect on anaerobic C02 evolution. Hydrogen oxidation was inhibited by both formulations in aerobic and anaerobic soil. Aerobic oxidation was inhibited to the same extent by both formulations; Roundup had a stronger inhibitory effect on anaerobic H2 oxidation than did glyphosate. No toxicity to any of these activities should be seen at recommended field application rates of the herbicide. . 1988. Glyphosate in the environment. Water, Air, and Soil Pollution 39 : 409-20. N-(phophonomethyl) glycine (glyphosate) is an extremely effective broad spectrum herbicide. This manuscript reviews glyphosate metabolism in plants and yeasts, its uses in agricultural applications, interactions with soil and water, glyphosate biodegradation, and effects on microbial activities and populations in soil.

2.

3.

4.

5.

127

Plant and Soil Residues 6. Cessna, A. J., A. L. Darwent, K. J. Kirkland, L. Townley-Smith, K. N. Harker, and L. P. Lefkovitch. 1994. Residues of glyphosate and its metabolite AMPA in wheat seed and foliage following preharvest applications. Canadian Journal of Plant Science 74 , no. 3 : 653-61 . In a 2-yr study at four locations in western Canada, residues of glyphosate and its major metabolite aminomethyl-phosphonic acid (AMPA) were measured in the seed and foliage of wheat (Triticum aestivum L.) following preharvest applications at rates of 0.45, 0.9 or 1.7 kg acid equivalent/ha. Herbicide treatments were applied in early August to mid-September at seed moisture contents ranging from 52 to 12%. Glyphosate and AMPA residues in the seed increased as the rate of application increased, and decreased as the seed moisture content at the time of application decreased. However, when the maximum application rate of 1.7 kg/ha was sprayed at seed moisture contents of 40% or less, glyphosate residues in the seed were lt 5 mg/kg, the Maximum Residue Level recently established by Health Canada. Glyphosate and AMPA residues in the straw also increased with increasing application rate, but there was no consistent pattern in residues of either chemical with seed moisture content at the time of application. Physiological maturity of the crop, rainfall washoff, and application rate appeared to play important roles in determining the magnitude of glyphosate and AMPA residues in the seed and straw of wheat. Cessna, A. J., and J. Waddington. 1995. Dissipation of glyphosate and its metabolite AMPA in established crested wheatgrass following spring application. Canadian Journal of Plant Science 75 , no. 3 : 759-62 . Using glyphosate to suppress resident vegetation while establishing sod-seeded alfalfa in pastures may result in unacceptably high residues within the treated foliage. In a 2-yr study, the dissipation of glyphosate and its major metabolite, AMPA, was measured in the foliage of established crested wheatgrass (Agropyron desertorum (Fisch.) Schult.) following spring application at 2.2 kg a.i./ha. Within 2 wk, glyphosate residues had decreased to lt 50 mg/kg, the international MRL for fodder of grasses. Washoff by rainfall appeared to be the major route of dissipation. Residues of AMPA were generally about one order of magnitude less than the corresponding glyphosate residues. Chizhov, B. E. 1986. Changes in the herbaceous shrub layer in green moss and pine forests caused by the chemical treatment of the soil for forest cultures. Lesovedenie 1: 58-66. Contemporary herbicides provide regulation of live ground cover in the process of forest recovery. Differences in spectra of phytotoxic action of dalapon, glyphosate, atrazine, gardoprim, velpar allow to choose such a preparation which could suppress undesirable plants and preserve valuable ones with regard for features of vegetation in felled areas. (Soviet Union). Chkanikov, D. I., L. V. Rimarenko, and A. M. Nazarova T. A. Makeev. 1987. Analysis of glyphosate residues in plants. Agrokhimiya 4: 117-23. Literature on the analysis of glyphosate residues is reviewed with reference to physicochemical properties, extraction methods, extract purification and methods of quantitative determination including GLC, HPLC and TLC. (Soviet Union). Clower, M. 1991. Pesticide residues in food. U.S. Food and Drug Administrations Program for immunoassay. ACS Symposium Series No. 451 pp. 49-58. The use of immunoassays as a means of testing for pesticide residues in foods are currently being assessed by the U.S. Food and Drug Administration (FDA). The potential benefits and considerations for the development of immunoassays for regulatory analyses, potential applications of enzyme immunoassay (EIA) methods in residue monitoring and FDA research on commercial immunoassay kits and EIA method development are discussed. Specific immunoassay methods that are being developed for the detection of paraquat in potatoes, fenamiphos and its sulfoxide and sulfone metabolites in oranges, carbendazim, benomyl and thiophanate-methyl in apples and glyphosate in soyabeans are also described. Comes, R. D., V. F. Bruns, and A. D. Kelley. 1976. Residues and persistence of glyphosate in irrigation water. Weed Science 24: 47-50. Neither glyphosate (N-(phosphonomethyl)glycine) nor the soil metabolite aminomethylphosphonic acid were detected in the first flow of water through the two canals following 128

7.

8.

9.

10.

11.

Plant and Soil Residues application of glyphosate at 5.6 kg per ha to ditchbanks when the canals were dry. Soil samples collected the day before canals were filled (about 23 weeks after treatment) contained about 0.35 ppm glyphosate and 0.78 ppm aminomethylphosphonic acid in the 0 to 10-cm layer. When glyphosate was metered into the water at a rate calculated to provide 150 ppb in the canal water at a single site on two flowing canals, about 70% of the glyphosate was accounted for 1.6 km downstream from the application site. Thereafter, the rate of disappearance diminished, and about 58% of the applied glyphosate was present at the end of the canals 8 or 14.4 km downstream from the introduction sites. 12. Cornish, P. S. 1992. Glyphosate residues in a sandy soil affect tomato transplants. Australian Journal of Experimental Agriculture 32: 395-99. (Australia). Coupland, D., and P. J. W. Lutman. 1982. Investigations into the movement of glyphosate from treated to adjacent untreated plants. Association of Applied Biologists 101: 315-21. Transfer of glyphosate from treated to adjacent untreated plants was investigated under glasshouse conditions using wheat and Agropyron repens. When glyphosate was used at concentrations characteristic of conventional field application rates, and where shoot contact was prevented, no symptoms were observed on untreated plants. When there was shoot contact, and when glyphosate was used at 2 kg a.i./ha (10 g/l), phytotoxic effects were observed on untreated plants. At higher concentrations of glyphosate (90 or 180 g a.i./l), typical of selective applications with ropewick or roller applications, evidence of root transfer of herbicide was found. In pot experiments these phytotoxic effects were variable due, perhaps, to variable amounts of root contact. Confining the roots, by growing the plants in tubes, increased the level of phytotoxicity. (United Kingdom). Couture, G., J. Legris, and R. Langevin. 1995. valuation des impacts du glyphosate utilis dans le milieu forestier, Qubec. Gouvernement du Qubec, ministre des Ressources naturelles, Direction de lenvironment forestier, Service du suivi environmental , Qubec. Publ. RN95-3082. 187 p. Glyphosate is a phytocide registered in Canada for use in forests since 1984. It has been in use in Qubec since 1985. In its commercial form (VisionTM), it is used mainly in release treatment of softwood regeneration, by means of broadcast aerial or ground applications. The ministre des Ressources naturelles has performed an environmental follow-up study of this phytocide in operational conditions. Glyphosate is also available in the form of an encapsulated water-soluble paste, sold under the trade names of EzjectTM and Gel CapTM-G. This form is administered by stem or stump injection. However, it is not widely used in Qubec. Cowel, J. E., J. L. Kunstman, P. J. Nor, J. R. Steinmetz, and G. R. Wilson. 1986. Validation of an analytical residue method for analysis of glyphosate and metabolite: An interlaboratory study. Journal of Agricultural and Food Chemistry 34: 955-60. A new residue method for the analysis of glyphosate and (aminomethyl)phosphonic acid has been validated with an interlaboratory study. Five different analysts from Monsanto Co. and other laboratories participated in testing of five different matrixes: alfalfa forage, cabbage, grapes, soybean grain, environmental water. These were chosen to represent the wide variety of matrixes analyzed for glyphosate-related residues. The cornerstone of the method is concentration and isolation via chelation ion exchange, with subsequent quantitation by HPLC with postcolumn reaction detection. The method was validated over the concentration range from 0.05 to 5.00 ppm with overall analytical recoveries of 80.9 13.8% for glyphosate and 79.2 13.8% for (aminomethyl)phosphonic acid. The coefficient of variation for both analyses was 17%, which fits well with that predicted for the analysis of compounds in this concentration range. Crisanto, T., M. J. Sanchez-Martin, M. Sanchez-Camazano, and M. Arienzo. 1994. Mobility of pesticides in soils. Influence of soil properties and pesticide structure. Toxicology and Environmental Chemistry 45, no. 1-2: 97-104. A comparative study was made of the mobility of five pesticides with different structures in 13 natural soils with different chemical and physical properties. Pesticide mobility was determined using soil thin layer chromatography and super (14)C-labeled compounds. According to the Rf values,

13.

14.

15.

16.

129

Plant and Soil Residues pesticide mobility varied in the following order: acephate > fluometuron > atrazine > ethofumesate > glyphosate, acephate being highly mobile, while atrazine, fluometuron and ethofumesate were mobile, and glyphosate was immobile. The results provided by linear regression analysis show highly significant correlation coefficients (p < 0.001) between the Rf values of atrazine, ethofumesate and fluometuron and soil organic matter content. There was also a significant correlation (p<0.01) of the Rf of glyphosate with the soil silt + clay content. Despite this, it was found that dissolved organic matter influences the mobility of the five pesticides and that the soil phosphate content affects the mobility of glyphosate. (Spain). 17. Darwent, A. L., K. J. Kirkland, L. Townley-Smith, K. N. Harker, A. J. Cessna, L.P. Lukow, and O.M. Lefkovitch. 1994. Effect of preharvest applications of glyphosate on the drying, yield and quality of wheat. Canadian Journal of Plant Science 74, no. 2: 221-30. In experiments conducted from 1988 to 1990 at 4 locations in the Parkland zone of western Canada, the drydown of seed and foliage, seed yield, seed quality and baking quality of wheat (Triticum aestivum L.) following preharvest applications of glyphosate were compared with those following windrowing prior to harvest or direct cutting of the standing crop. Glyphosate was applied in late July to early Sept. at rates of 0.45, 0.9 and 1.7 kg acid equivalent/ha to wheat with seed moisture contents ranging from 80 to 11%. There was little or no difference in 1000-seed wt., sample density, seed germination and protein content from plots sprayed with glyphosate at seed moisture contents <40% than from control plots windrowed at the same moisture content or direct cut at maturity. Baking quality was not affected by any of the glyphosate treatments in experiments where the wheat was harvested by direct cutting. Devlin, R. M., S. J. Karczmarczyk, I. I. Zbiec, and Z. K. Koszanski. 1986. Initial and residual activity of glyphosate and SC-0224 in a sandy soil. Crop Protection 5: 293-96. Initial and residual herbicide activity of glyphosate and SC-0224 (trimethylsulphonium carboxymethylaminomethylphosphonate) when applied to a sandy soil was investigated. A bioassay employing wheat (Triticum vulgare L., cv. mericopa) was used to determine the residual activity of the herbicides on different characteristics of plant growth. At 5 kg/ha both herbicides significantly reduced shoot length. This was observed in wheat planted immediately as well as 10 days after application. SC-0224 was more active than glyphosate, significantly reducing shoot length at the 2.5 kg/ha rate 10 days after application. Root length was reduced by both herbicides at 2.5 and 5 kg/ha in wheat planted 10 days after application. Reductions in fresh and dry weight were also observed for both shoots and roots. It is apparent that SC-0224 and glyphosate have residual herbicide activity in sandy soil, at least up to 10 days after application, and that SC-0224 has more herbicidal activity than glyphosate in this respect. Dostie, R. 1993. Dpt de glyphosate a l'extrieur des aires traites en 1991. Direction de l'environment, Service du suivi environmental, Government of Quebec. FP93-3117. 17 p. Each year, the Direction de l'environment of the Ministry of Forests (MFO) conducts environmental follow-up of glyphosate sprayings in the forest environment. In 1991, to validate the width of buffer strips along sensitive zones, it was deemed essential to evaluate glyphosate deposits outside the treated areas. The width of buffer strips had been determined according to the quantities of glyphosate collected during experimental sprayings and according to thresholds of 10 g/ha and 100 g/ha, which were deemed safe for adequate protection of sensitive zones. These thresholds were fixed on the basis of glyphosate data. The treatments were applied from a plane, a helicopter, a skidder equipped with a drum and a carrier equipped with ramps. Shortly before spraying, pieces of aluminum foil were set out on stakes placed at different distances from the areas to be treated. Gas chromatography was used to quantify the glyphosate deposited on these artificial collectors. After spraying the drum-equipped skidder, the maximum quantities of glyphosate detected were 1.55 g/ha and 0.71 g/ha at 10 m and 25 m, respectively, from the treated areas. With the ramp-equipped carrier, glyphosate quantities did not exceed 4.29 g/ha at >5 m. For areas sprayed from the helicopter and the AgCat and Cesna airplanes, the maximum quantities of glyphosate at 60 m were 35 g/ha, 116 g/ha and 8 g/ha, respectively. At 100 m, the maximum quantities detected were 28 g/ha, 59 g/ha and 7 g/ha. In relation to the applied dose of 1495 g/ha of glyphosate, these values represent less than 8% at 60 m and less than 4% at 100 m from the treated areas. Considering the thresholds of 100 g/ha of

18.

19.

130

Plant and Soil Residues glyphosate at 10 m and 10 g/ha at 25 m, the width of the buffer strips determined for ground-level spraying is sufficient to adequately preserve sensitive zones. It even provides an appreciable safety factor. The 60-meter buffer strips required for aerial glyphosate sprayings, with a tolerance of 100 g/ha, is also adequate. However, the maximum quantity of 10 g/ha tolerated at 100 m was exceeded during a helicopter-applied treatment and the two treatments made with the AgCat airplane. According to the results, only the threshold of 10 g/ha of glyphosate set out for the 100-meter buffer strip was exceeded, and this only during certain aerial sprayings. Observations made during those spraying showed that quantities of glyphosate exceed the threshold mainly when procedures do not conform to the standards set out by the MFO. Signalling, verification of weather conditions and monitoring of operations should be conducted each time there is aerial spraying. The next evaluations of glyphosate deposits outside treated areas should mainly be carried out during aerial spraying. It is important to obtain more information about glyphosate deposits at a distance of 100 m from treated areas. 20. Edwards, W. M., G. B. Triplett, and R. M. Kramer. 1980. A watershed study of glyphosate transport runoff. Journal of Environmental Quality 9: 661-65. Glyphosate (N-phosphonomethyl glycine) formulated as Roundup herbicide, was applied on 0.3- to 3.1-ha watersheds at rates of 1.10-, 3.36-, and 8.96-kg/ha as a preseeding herbicide in the notillage establishment of fescue and corn. Runoff from natural rainfall following early springtime treatment was measured and analyzed to define concentration and transport of glyphosate under these conditions. The highest concentration of glyphosate (5.2 mg/l) was found in runoff occurring 1 day after treatment at the highest rate. Glyphosate (2 g/l) was detected in runoff from this watershed up to 4 months after treatment. For the lowest application rates, maximum concentration of the herbicide in runoff was less than 100 g/l for events occurring 9-10 days after application and decreased to less than 2 g/l within 2 months of treatment. The maximum amount of glyphosate transport by runoff was 1.85% of the amount applied, most of which occurred during a single storm on the day after application. In each of the three study years, herbicide transport in the first runoff event following treatment accounted for 99% of the total runoff transport on one watershed. Glyphosate residues in the upper 2.5 cm of treated soil decreased logarithmically with the logarithm of time; they persisted several weeks longer than in the runoff water. Ernst, W. R., R. Morash, W. Freedman, and K. Fletcher. 1987. "Canada Surveillance Report ." Measurement of the environmental effects associated with forestry use of Roundup. EP-5-AR87-8. A number of studies were undertaken to determine the environmental effects of RoundupR and, in some instances, to compare these effects with those documented for 2,4-D and 2,4,5-T. The deposit of ground and aerially applied glyphosate on flat plates was measured using rhodamine dye as a tracer. Applications were at normal and twice normal operational dosage rates. Deposit was highly variable. However, ground applications provided the highest rate of deposit inside the target area with no measurable off-site deposits. Aerial sprays resulted in measurable deposit up to 95 metres off target. The persistence of glyphosate on foliage and litter from aerially sprayed plots was also measured. Foliage selected for study were from red maple (Acer rubrum), balsam fir (Abies balsamea), raspberry (Rubus spp.) and grasses (Poaceae and Cyperaceae). Highest concentrations of glyphosate were detected the day after spraying in raspberry and grasses (13 and 12 g/g dry weight, respectively). Red maple, forest litter and balsam fir had smaller concentrations of 5.1, 2.6 and 0.09 g/g dry weight, respectively, one day after spray. Residues in all substrates, with the exception of forest litter, declined rapidly up to the last day of sampling, 58 days post spray. Forest litter residues demonstrated an overall decline, however, the rate was less rapid than in foliage. The rate of decomposition of litter, both in the field and in the laboratory, after treatment with herbicide was also explored. Substrates studied were forest floor material (litter, duff and humus), and fresh foliage from red maple or white spruce (Picea glauca). Samples of substrate were either dosed in the laboratory or placed in field plots which were subsequently sprayed with herbicide. The amount of substrate decomposition decreased with increasing dosages of 2,4-D, 2,4,5-T, a 50:50 mixture of 2,4-D and 2,4,5-T and glyphosate. These effects, however, were only significant at concentrations greater than 1000 g/g. Such concentrations are considerably in excess of those associated with operational spraying. The final study determined stream residue concentrations following direct oversprays at operational dosage rates of glyphosate and evaluated whether such concentrations affect stream periphyton abundance. Artificial substrates were placed in a stream above and below the target area, sampled at various times after spray and 131

21.

Plant and Soil Residues analyzed for chlorophyll a, phaeophytin, biomass and ATP content. Stream residues were highly variable immediately after spray with the highest concentration (39 g/l) being measured 30 hours post spray. While periphyton abundance was variable, no changes which could be attributed to pesticide effects were noted for the study period which concluded 37 days after spray. 22. Feng, J. C., and D. G. Thompson. 1989. Persistence and dissipation of glyphosate in foliage and soils of a Canadian coastal forest watershed. Proceedings of the Carnation Creek Herbicide Workshop.: Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 65-87. The environmental persistence and behaviour of glyphosate and its major metabolite, aminomethylphosphonic acid (AMPA), were investigated in terrestrial components of a Canadian coastal forest watershed, following aerial application of the herbicide Roundup. Soil samples were analyzed to determine persistence and leaching of glyphosate and AMPA in both seasonally flooded and well-drained soils. Foliage and leaf litter of salmonberry (Rubus spectabilis) and red alder (Alnus rubra Bong) were monitored to determine initial foliage residue, residues associated with leaf litter and to estimate the degree of residue input via leaf fall. Average estimates of initial deposit rate were close to the nominal application rate of 2.0 kg/ha. Deposition rates varied between sites, ranging from 0.60 kg/ha to 3.42 kg/ha. Analyses of initial foliage samples also indicated highly variable chemical deposition. Residue levels in both salmonberry and leaf litter collected 15 days post-application were significantly lower than initial foliage residues. Leaf litter residues of glyphosate and AMPA in both species declined over time, with less than 1% remaining 29 days post-application. Soil residue determinations showed that both glyphosate and AMPA were retained primarily in the upper, organicrich soil layer, 0-15 cm. Residues of glyphosate declined with time at all locations, with only 6-18% of initial levels remaining after 360 days. The results indicate that glyphosate and AMPA are nonpersistent in terrestrial components of the watershed and that neither glyphosate nor AMPA are susceptible to leaching under the conditions studied. There was no evidence of potential for long-term groundwater or surface water contamination. . 1990. Fate of glyphosate in a Canadian forest watershed. 2. Persistence in foliage and soils. Journal of Agricultural Food and Chemistry 38: 1118-25. Residues of glyphosate [N-(phosphonomethyl)glycine] and the metabolite (aminomethyl)phosphonic acid (AMPA) were monitored in foliage, leaf litter, and soils following aerial application of Roundup herbicide (nominal rate 2.0 kg/ha AI) to the Carnation Creek watershed of Vancouver Island, British Columbia. Glyphosate deposit was variable, ranging from 1.85 to 2.52 kg/ha AI, depending upon location within the watershed. Foliar residues in red alder and salmonberry were 261.0 and 447.6 g/g, respectively, indicating good impingement on the target foliage. Leaf litter residues, which averaged 12.5 g/g for red alder and 19.2 g/g for salmonberry initially, declined to less than 1 g/g within 45 days postapplication (DT50 < 14 days). In soils, glyphosate and AMPA residues were retained primarily in the upper organic layers of the profile, with > 90% of total glyphosate residue in the 0-15-cm layer. Distribution data for both glyphosate and AMPA suggested strong adsorption and a low propensity for leaching. Glyphosate soil residues dissipated as a function of time and an estimated DT50 of 45-60 days. After 360 days, total soil residues of glyphosate were 618% of initial levels. Feng, J. C., D. G. Thompson, and P. E. Reynolds. 1989. Fate of glyphosate in a forest stream ecosystem. Proceedings of the Carnation Creek Herbicide Workshop, Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 4564. See Water Quality Section. Fletcher, K., and B. Freedman. 1986. Effects of the herbicides glyphosate, 2,4,5-T and 2,4-D on forest litter decomposition. Canadian Journal of Forest Research 16: 6-9. Laboratory studies with two leaf litter and one forest floor substrate showed that the herbicides 2,4-dichlorophenoxyacetic acid (2,4-D), 2,4,5-trichlorophenoxyacetic acid (2,4,5-T), a 50:50 mixture of these, and glyphosate all had toxic thresholds at which they reduced decomposition. However, in all cases, the thresholds were > 50 times higher than residue concentrations that occur in the field after 132

23.

24.

25.

Plant and Soil Residues silvicultural herbicide treatments. In a field study at one site, no measurable 1-year postspray effects on litter decomposition were found among treatment plots sprayed at 0.0, 3.4 or 6.7 kg 2,4,5-T/ha. 26. Fujiie, A., T. Yokoyama, M. Fujikata, M. Sawada, and M. Hasegawa. 1993. Pathogenicity of an entomogenous nematode Steinernema kushidai Mamiya (Nematode: Steinemematidae), on Anomala cuprea (Coleoptera: Scarabaeidae). Japanese Journal of Applied Entomology and Zoology 37: 53-60. See Terrestrial Invertebrates Section. Garabito, G. E. A., Y. Sandoval, and A. H. Trujillo. 1991. Effect of some herbicidal compounds used in tropic areas on the microbial populations in soil. Reviews of Latinoam Microbiology 33: 191-201. The effects that some herbicidal compounds have on soil microbial populations were studied. The agrochemicals studied were: Glyphosate, Paraquat, Ametrine, 2,4-D, Atrazine and Atrazine + Terbutrine. The effect was measured by plate counts of bacteria, actinomycetes and fungi on three types of soils (Feozem, Castanozem and Chernozem) from Tecoman Col. Mexico, during nine weeks after treatment with agrochemicals in both field and laboratory conditions. Results showed that microbial populations tend to normality around 6 weeks after the exposure to agrochemicals. (Mexico). Gerritse, R. G., J. Beltran, and F. Hernandes. 1996. Adsorption of atrazine, simazine, and glyphosate in soils of the Gnangara Mound, Western Australia. Australian Journal of Soil Research 34, no. 4: 599-607. Sandy soils were sampled from second rotation sites of Pinus pinaster Ait. on the Gnangara Mound in Western Australia. Adsorption isotherms were measured for atrazine (6-chloro-N-2-ethyl-N4-isopropyl-1,3,5-triazine-2,4-diamine), simazine (6-chloro-N,N'-diethyl-1,3,5-triazine-2,4-diamine), and glyphosate (N-phosphonomethyl-aminoacetic acid). Adsorption isotherms were also measured for degradation products of atrazine: hydroxyatrazine (6-hydroxy-N-2-ethyl-N-4-isopropyl-1,3,5-triazine-2,4diamine) (HA), desethylatrazine (6-chloro-N-isopropyl-1,3,5-triazine-2,4-diamine) (DEA); and of glyphosate: aminomethylphosphonic acid (AMPA). The adsorption of the 2 triazines was proportional to soil organic carbon content and was not affected significantly by other soil parameters. The affinity for soil organic carbon was in the order atrazine = simazine = DEA gt HA. Affinity of atrazine for the type of organic matter in the Gnangara Mound soils (expressed as K-oc) was significantly greater than is commonly reported for other soils. The adsorption of glyphosate and AMPA increased strongly with iron and aluminium content of soils and decreased with increasing soil organic carbon content. This would indicate that glyphosate and AMPA are mainly adsorbed by clay minerals, while soil organic matter competes for adsorption sites and inhibits adsorption. Contrary to what is usually reported for batch adsorption of pesticides in soils, significant increases in adsorption of the triazines and glyphosate were measured after 1 day of equilibration. (Australia). Gianfreda, L., F. Sannino, N. Ortega, and P. Nannipieri. 1994. Activity of free and immobilized urease in soil: Effects of pesticides. Soil Biology and Biochemistry 26, no. 6: 777-84. The effects of atrazine, carbaryl, glyphosate and paraquat on the activity of jack bean urease, free in solution or immobilized on a clean clay (montmorillonite), an organic compound (tannic acid) or a synthetic organo-mineral (A l(OH)-x-tannate and Al (OH)-x-tanate-montmorillonite) complexes were studied. For comparison, pesticide influence on urease activity in some soils and soil extracts was investigated, under laboratory conditions. Urease was affected by pesticides to different degrees, depending on its physical state, i.e. if free, immobilized or as total activity in soils or soil extracts. Glyphosate and paraquat enhanced the urease activity of soils (by 1.1-1.4-fold) and soil extracts (by 2.59-6.73-fold). No significant effects were detected on the activity of jack bean urease, either free in solution or absorbed on montmorillonite. Similarly, no influence on the activity of urease immobilized on synthetic organo-mineral complexes was observed with paraquat. Since atrazine and carbaryl, which have low solubility in water, were applied in methanol, the effect of this solvent on urease activity was investigated. The activity of urease in soils and soil extracts was strongly increased by methanol (by 1.48-7.47-fold and by 1.42-11.62-fold, respectively) and carbaryl (by 1.73-2.91-fold). Atrazine partly reduced the increasing effect of methanol. Free and immobilized urease were generally inhibited

27.

28.

29.

133

Plant and Soil Residues by methanol and both pesticides (on average -40%), but the extent of inhibition greatly depended on the nature of immobilizing support. (Italy). 30. Gianfreda, L., F. Sannino, and A. Violante. 1995. Pesticide effects on the activity of free, immobilized and soil invertase. Soil Biology and Biochemistry 27, no. 9: 1201-8. The influence of four pesticides (atrazine, carbaryl, glyphosate and paraquat) on the catalytic behaviour of invertase, either free, immobilized on inorganic and organic soil colloids or in soils, was investigated. Invertase was immobilized on a clean clay (montmorillonite), an organic compound (tannic acid), and on synthetic organo-mineral (Al(OH)-x-tannate and Al (OH)-x-tanate-montmorillonite) complexes. Soils with different physico-chemical properties were utilized. The effects of pesticides on invertase performance depended not only on the nature of the pesticide but also on the status or the enzyme, i.e. if free, immobilized or in soil. Glyphosate and paraquat enhanced the activity of invertase either free or immobilized on montmorillonite and both pesticides behaved as mixed-type non-essential activators. Activity decreases were instead measured for the enzyme immobilized on organic and organo-mineral matrices. Contrasting results (increases, decreases and no effects) were detected for soil invertase. A general inhibition effect was exhibited by methanol on free, immobilized or soil invertase, but the extent of inhibition depended on the enzyme microenvironment. The addition of atrazine and carbaryl caused partial increases of free and immobilized invertase activity, whereas carbaryl further reduced enzymatic activity in some soils. (Italy). Glass, R. L. 1987. Adsorption of glyphosate by soils and clay minerals. Journal of Agricultural and Food Chemistry 35: 497-500. The adsorption of glyphosate by several soils and clay minerals was investigated with a highperformance liquid chromatographic method. Freundlich adsorptive capacities (K) were determined as 138, 115, and 8, respectively, for the clay minerals montmorillonite, illite, and kaolinite; and 76, 56, and 33, respectively, for Houston clay loam, Muskingum silt loam, and Sassafras sandy loam. Adsorption by the soils appeared to be related to the clay content and the cation-exchange capacities of the soils. Adsorption of glyphosate by illite was found to be independent of pH. The loss of solution protons as revealed by the increase in pH of the glyphosate solutions equilibrated with illite suggested the existence of a cation-exchange reaction. Saturating montmorillonite with various cations increased glyphosate adsorption in the order Na+<Ca+2<Mg+2<Cu+2<Fe+3. The complexation of glyphosate by cations released from cation-saturated clays via a cation-exchange reaction with solution protons was proposed as an adsorption mechanism. The formation of a new absorbance band (.Lambda. Max =226 nm) in the UV spectra of glyphosate solutions (pH4) equilibrated with Cu+2-montmorillonite provided additional evidence for complex formation between glyphosate and cations. Goldsborough, L. G., and D. J. Brown. 1993. Dissipation of glyphosate and aminomethylphosphonic acid in water and sediments of boreal forest ponds. Environmental Toxicology and Chemistry 12: 1139-47. See Water Quality Section. Gutierrez, E., F. Arreguin, R. Huerto, and P. Saldana. 1994. Aquatic weed control. International Journal of Water Resource Development 10, no. 3: 291-312. A general discussion is provided of the water hyacinth (Eichhornia crassipes) which may become a weed as a result of its uncontrolled reproduction, and of the five main control methods. The Mexican Aquatic Weed Control Program (PROCMA) is described, including its goals and guidelines. Information is given concerning COD, DO, pesticide residues, pH, conductivity and total phosphorus resulting from the use of 2,4-D, diquat and glyphosate in three dams infested with water hyacinth and cat-tail. (Mexico). Hance, R. J. 1976. Adsorption of glyphosate by soils. Pesticide Science 7: 363-66. The adsorption of glyphosate in a slurry system by 9 soils was correlated with an arbitrary single point inorganic phosphate sorption index which gave a measure of unoccupied phosphate sorption capacity, but not with total phosphate sorption capacity at pH 2.6, the level of iron and aluminium soluble in Tamm's acid oxalate or with clay content, organic matter content or pH. It is concluded that inorganic phosphate excludes glyphosate from sorption sites. The extent of adsorption 134

31.

32.

33.

34.

Plant and Soil Residues was higher than, though of the same order as, that of diuron though the two are not related. From consideration of the results of previous experiments with plants grown in culture solution it is suggested that the low phytotoxicity of glyphosate applied to the soil is not due simply to high adsorption. Rather it is the result of a combination of moderate adsorption and the low intrinsic activity of this compound when made available to the root system. 35. Heinonen-Tanski, H. 1989. The effect of temperature and liming on the degradation of glyphosate in two arctic forest soils. Soil Biology and Biochemistry 21: 313-18. Degradation of C14-labelled glyphosate was studied in two arctic forest soils under simulated autumn and spring temperatures prevalent for the region. A rapid initial degradation of glyphosate, as measured by C14O2 evolution, was observed in a Hylocomium-Myrtillus type (HMT) soil which was significantly higher than in a Myrtillus-Calluna-Cladina type-Empetrum-Myrtillus type (MCCIT-EMT) soil at an initial soil temperature of 15oC. Degradation was found to be temperature dependent in both soils, over a range of 1.9-15oC. A decrease of 10oC reduced the degradation to one-tenth and onefifteenth of the initial rate in the MCCIT-EMT and HMT soils, respectively. At the end of the temperature simulated incubation (217 days) high levels of radioactivity were detected in a range of soil fractions studied. Liming of both soils resulted in a slight initial increase in degradation rate for the HMT soil, but throughout the incubation the effect of liming was not significant in either soil. (Finland). Heinonen-Tanski, H., L. Montonen, L. R. Ervio, S. Junnila, and B. Pessala. 1985. The effects of chlorsulfuron, glyphosate, metribuzin and TCA on soil nitrification capacity and dehydrogenase activity. In: Comportement et Effects Secondaires des Pesticides dans le Sol. Versailles, France. Pp. 183-189. In field trials at Jokioinen and Ruukki, Finland, potatoes were treated with glyphosate, metribuzin or TCA and barley with chlorsulfuron all applied at the normal rate or 3 times this rate. The effects of the herbicides on soil dehydrogenase and nitrification activity were studied. Chlorsulfuron and glyphosate had little effect on soil dehydrogenase and nitrification, but metribuzin and TCA reduced both activities especially in clay soils. The effects of TCA were still evident 1 year after application. (Finland). Heinonen-Tanski, H., C. Rosenberg, H. Siltanen, S. Kilpi, and P. Simojoki. 1985. The effect of the annual use of pesticides on soil microorganisms, pesticide residues in the soil and barley yields. Pesticide Science 16: 341-48. A study has been made of the influence of pesticides used annually on soil microorganisms and crop yields (Finland). The persistence of these pesticides in the soil was also investigated. The herbicides MCPA, glyphosate, maleic hydrazide and triallate, and the insecticide parathion, were applied on experimental plots on which barley was grown during the years 1973-1981. The fungicide 2methoxyethylmercury chloride was used every year for dressing the seeds grown in pesticidetreated plots. The pesticide treatments did not affect significantly the numbers of several groups of soil microorganisms. A slight increase was, however, observed in the nitrification activity in the soil. The barley yields were on average higher on pesticidetreated plots than on controls because of successful weed control. Pesticide residues in the soil were generally very low; for example, for parathion they were below 0.02 mg/kg within 11 days, and for MCPA 0.06 mg/kg within 7 days. However, the glyphosate residue was 1.6 mg/kg in the autumn 2 days after the treatment, and the residue settled to a level of 0.2 mg/kg during the following summer. No clear dependence was observed between the residue level and the time between treatment and sampling. (Finland). Helweg, A. 1986. Side effects caused by pesticide combinations. FEMS (Federation of European Microbiological Societies) Symposium No. 33. Microbial Communities in Soil. V. Jensen, A. Kjoller, and L. H. Sorensen (editors). (Denmark). Ismail, B. S., A. J. Kader, and O. Omar. 1995. Effects of glyphosate on cellulose decomposition in two soils. Folia Microbiologica 40, no. 5: 499-502. Glyphosate with an equivalent concentration of either 0, 2.16 or 8.64 kg/hm-2 was sprayed on to cellulosic materials before burying in two soil types; peat (soil I) and sandy clay loam (soil II). 135

36.

37.

38.

39.

Plant and Soil Residues Alternatively the soils were sprayed with 0, 20 or 150 ppm of the herbicide before burying the cellulosic material either immediately or after preincubation for 4 weeks. In soil I, the increase in glyphosate concentrations subsequently reduced the decomposition of cellulosic material regardless of the method of application employed. Glyphosate at 8.64 kg/hm-2 reduced the mass loss of the treated substrate by 83%. However, cellulose decomposition in soil preincubated for 4 weeks before burying was affected almost to the same extent as the untreated control. Glyphosate stimulated cellulose decomposition when substrates were buried in soil II. Mass loss in soil treated with 150 ppm increased by about 100% while when glyphosate was sprayed directly to the substrate (at 8.64 kg/hm-2), the loss was about 25%. (Malaysia). 40. Judy, B. M., W. R. Lower, C. D. Miles, M. W. Thomas, and G. F. Krause. 1990. Chlorophyll fluorescence of a higher plant as an assay for toxicity assessment of soil and water. American Society for Testing and Materials. Symposium on use of Plants for Toxicity Assessment W. Wang, J. W. Gorsuch, and W. R. Lower (editors), pp. 308-18. Junnila, S. 1986. An overview of trials concerning mobility and transport of herbicides. Nordisk Jordbrugsforskning 68: 384-85. This paper was presented at the Nordiske Jordbrugsforskeres Foreming Workshop on leaching of chemical control compounds, held in Uppsala in Nov. 1985. In field trials at the Agricultural Research Centre, Jokioinen, and at Norra-Osterbotten starting in 1982, the persistence and mobility of 6 herbicides and their effect on soil microorganisms were investigated. Glyphosate was applied in autumn and TCA in spring and autumn before potatoes were planted. Metribuzin and flurochloridone were applied to potatoes in June. In spring cereals, chlorsulfuron and metsulfuron-methyl were tested at 4 (normal rate) and 12 g/ha. Soil samples were taken at 0-5, 5-15 and 15-25 cm depths, 1 day and 1 month after spraying, and also at the beginning and end of the growth period, so long as test plants in greenhouses showed that residues were still present. Trials were on sand, loam and peat soils at Jokioinen and on sand and peat soils at Norra-Osterbotten. In sandy soils TCA moved quickly to a depth of 25 cm and disappeared after 6 months. In loam, TCA was more readily bound, and residues were found after a year at the normal rate and after 2.5 years at the higher rate. TCA and chlorsulfuron tended to move vertically through the soil. With spring application, TCA reached a depth of 25 cm after 1 month. Glyphosate was strongly adsorbed by soil. Ryegrass (Lolium sp.) test plants showed damage 6 months after glyphosate had been applied to the loam and sand soils and 1 year after application to peat soils in 1982. After spraying in 1983, plants were still damaged 1.5 years later. Metribuzin residues were found more than or equal to 1.5 years after application at 0.75 kg/ha and 2 years after application at 1.4 kg/ha. In 1983 metribuzin reached a depth of 25 cm in less than a month in sand and peat and by autumn in loam. In 1984, when rain during spraying was heavy, it reached a depth of 15 cm in 1 day. Flurochlorindone was found at a depth of 25 cm 1 day after sowing after abundant rain in 1984. At normal rates, residues remained for at least 1 growing season. Chlorosulfuron and metsufuron-methyl were both adsorbed quickly by all soil types and residues remained for at least one growing season but often for 1 year, especially at 3 times the normal rate. Soil tended to adsorb metsulfuron-methyl more than chlorsulfuron. (Sweden). . 1987. Transport and persistence of phytotoxicity of herbicides in organic soils. Sveriges Landbrukeuniversitet 37: 17-27. In field trials in Finland in organic soil, plots were treated with glyphosate, metribuzin, chlorosulfuron or metsulfuron-methyl at recommended and 3-fold rates. Soil samples were collected from 0-5, 5-15 and 15-25 cm depths. Persistence of glyphosate was low in organic soils; residues of the other herbicides occurred mainly in the top soil layers. Herbicides applied at the recommended rate caused no significant phytotoxicity to Lolium sp. and tomatoes with glyphosate, Lolium sp. and onions with metribuzin and onions with the other herbicides studied. Metsulfuron-methyl and chlorsulfuron caused significant damage to test plants even one year after treatment. (Finland).

41.

42.

136

Plant and Soil Residues 43. Kataoka, H., K. Horii, and M. Makita. 1991. Determination of the herbicide glyphosate and its metabolite aminomethylphosphonic acid by gas chromatography with flame photometric detection. Agricultural Biology and Chemistry 55: 195-98. A rapid, selective and sensitive method was developed for determining glyphosate and (aminomethyl)phosphonic acid (AMPA) by gas chromatography (GC). These compounds were converted into their N-isobutoxycarbonyl methyl ester derivatives and measured by GC with flame photometric detection, using fused-silica capillary column with cross-linked DB-1701. The calibration curves for glyphosate and AMPA in the range of 5-200 ng were linear, and the detection limits were about 10 and 15 pg as injection amounts, respectively. This method was applied to water and soil samples without a preliminary clean-up procedure, and glyphosate and AMPA were measured without any influence from coexisting substances. The recoveries of these compounds in potable water and river water samples were 96.2-100.3% and those in soil samples were 81.7-99.1%. (Japan). Khan, S. U., and J. C. Young. 1977. N-nitrosamine formation in soil from the herbicide glyphosate. Journal of Agricultural and Food Chemistry 25: 1430-1432. Formation of N-nitrosoglyphosate was observed when different soils were treated with sodium nitrite and the herbicide glyphosate at elevated levels. The highest yield was noted in soil low in organic matter and clay contents; however, nitrosation was not affected by soil pH. At low levels of glyphosate (5 ppm) and nitrite nitrogen (2 ppm) the formation of N-nitrosoglyphosate in soil was not observed. Korol, R. V. 1985. Regulation of glyphosate content in the soil. Gigiena I Sanitariya 8: 71-72. (Soviet Union). Kruglov, Y. V., N. B. Gersh, and M. Shtal'berg. 1980. The influence of glyphosate on the soil microflora. Khimiya v Sel'Skom Khozayistive 18: 42-44. (Soviet Union) . Laberge, L, J. Legris, and G. Couture. 1997. Glyphosate residues in pollen and honey after applications in an agro-forest environment. Draft Report . Ministre des Ressources naturelles du Qubec, Direction de lenvironement forestier Qubec . Glyphosate is a herbicide used by foresters to release regeneration from weed competition. In 1990, a study on glyphosate residues susceptible of being found in the honey of hives situated in the proximity to spraying areas was conducted in Quebec by the Ministry of Natural Resources (MNR). No glyphosate residues were detected (<0.05 g/g) in the seven samples analyzed. However, during the extraction phase of the honey, the combs of many apiaries are integrated together in the tank, which can lead to an important dilution of the glyphosate possibly present in the hives targeted by the sampling. Recommendations were made in order to analyze pollen samples that should have presented a dilution factor inferior to that of the honey, because the harvest can be performed daily. A preliminary study was carried out in 1991 with the collaboration of a pollen producer who owned apiaries near areas of glyphosate spraying. Two of the nine pollen samples showed detectable glyphosate residues of 1.62 and 0.14 g/g of fresh weight. Also, glyphosate (0.28 g/g) was detected in one of ten commercialized pollen samples. The following year, in 1992, a more detailed study was carried out in an agro-forest environment by the MNR in collaboration with the Quebec Ministry of Agriculture, Fish and Food (MAPAQ). Just before treatment, hives were placed inside the spraying sector (0 km), as well as 0.5 and 1 km away for three different locations (9 hives). The hives inside the treated areas were sheltered during the spraying. Pollen samples were collected twice a week for the first month following the herbicide treatment. Also, one sample of honey was collected before and after the pollen sampling period. Results (detection limit of 0.05 g/g) indicated that three days after treatment a maximum residues level of 8.2 g/g was detected in the pollen of a hive situated inside a treated area. Approximately 50% (39% - 61%) of the samples from the three locations contained glyphosate. Generally, residues declined with time after treatment and distance from treated areas. Three of the nine honey samples gathered after treatment showed detectable (>0.05 g/g) residues of glyphosate. Two of these samples were from treated areas. However, some control samples (honey and pollen) harvested before the application contained glyphosate indicating that the 137

44.

45. 46.

47.

Plant and Soil Residues residues were not the unique result of forestry use of glyphosate. Concurrently to the glyphosate study, for pollen and honey, 19 different pesticides were analyzed in the 9 control samples of the study and in 9 of after treatment samples. Moreover, 19 pollen samples and 13 honey samples of commercialized products were analysed for these pesticides. Results indicated the presence of many pesticides, mainly in pollen, in both commercialized and study samples. Finally, the MAPAQ made an evaluation of the toxicological risks associated with glyphosate and the other detected pesticides at these residue levels and concluded that the risks were negligible. 48. Legris, J., and G. Couture. 1989. Rsidues de glyphosate dans le feuillage, les ramilles et les fruits sauvages suite des pulvrisations terrestres en milieu forestier en 1985 et 1986. Government of Quebec, Ministry of Energy and Resources. Environmental Studies Service Publ. 3321. 27 p. In 1985 and 1986, the Ministry of Energy and Resources conducted an environmental study of glyphosate (RoundupTM). The herbicide glyphosate is used in ground applications for the upkeep of softwood plantations in public forests. The goal of this environmental monitoring was to determine the residual quantity and persistence of glyphosate in the foliage, twigs and wild fruit of the major ligneous species found in spraying areas. Near 35 samples, generally composite, were analyzed. These surfaces were treated with 1.5 kg of active ingredients per hectare. In foliage, the maximum concentration of residues was 45.3 g/g (fresh weight) in raspberry, 6 days following the treatment. After 27 days, 15% of this quantity was recorded and after 48 days, 6% was found. In birch twigs, the maximum quantity of glyphosate was 9.88 g/g, 3 days following the application. After 49 days, only 1.30 g/g was found, or about 13% of the maximum recorded. In wild fruit, results varied. The maximum residues detected were found in raspberries 7 days after the application and corresponded to 36.5 g/g. After 27 days, the value obtained was 0.139 g/g. In blueberries, residues showed 7.9 g/g after one day and dropped to 2.09 g/g after 22 days. Even considering the highest level of residues found in the vegetation and in fruit, there is little probability that wildlife and humans will suffer any significant direct consequences. However, residues in fruit have consistently been higher than the maximum limit established by the federal Food and Drug Act for chemical products in food (0.1 ppm) and it is the reason why panels were installed to forbid fruit picking. It is recommended that such samplings continue to be carried out to confirm and validate the results already obtained. . 1990. Rsidus de glyphosate dans un cosystme forestier suite des pulvrisations ariennes au Qubec en 1987. ER90-3085. Ministry of Energy and Resources, Environmental Studies Services. 35 p. In 1987, the Service of Environmental Studies of the Ministry of Energy and Resources conducted environmental monitoring of glyphosate levels occurring after aerial spraying for conifer release. The goal of this study was to determine the residual quantity of this phytocide in various components of the forest ecosystem. Three sectors located in the Bas-St-Laurent - Gaspsie region were treated using a helicopter equipped with a boom. The theoretical rate applied was 1.5 kg of active ingredient per hectare, but a study of deposits within the treated area revealed that only about 65% of this dose was actually present in the sectors covered. Sixty-one water samples were taken from five streams draining the sprayed areas during the first few days following treatment. The sampling points were located along the perimeter and 1 km from the treated areas. Along these watercourses, a buffer zone 60 meters wide was left unsprayed. Nine of these samples showed detectable residues (> 2 g/l). The highest residue level was 11 g/l observed one hour after treatment in proximity to the spraying site. The presence of residues in streams soon after spraying is probably due to a smaller droplet spectrum than expected, tending to cause drift. The presence of the phytocide in these environments at later periods may be due to precipitation occurring in the treated sectors before the sampling periods. To determine residue levels in a lentic environment, water samples were taken from three ponds directly exposed to spraying, one in each sector, 1 hour after spraying, then at 12-hour intervals over the next four days. A final sample was taken after 1 week. The highest residue level detected was 439 g/l, one hour after treatment. After 48 hours, concentrations were below 63 g/l and had declined to less than 17 g/l after one week. Thirty-one sediment samples were also collected from these ponds over a period of 27 months after spraying. The highest concentration observed was 1.58

49.

138

Plant and Soil Residues g/g dry weight after 21 months. Residue behavior in the sediments was variable from pond to pond, with values of 0.106, 1.45 and < 0.050 g/g after 27 months. In soil exposed to spraying, 5 samples composed of the 0-10 cm fraction were taken over the first 12 weeks following treatment. Only one value (0.151 g/g dry weight) was detected in excess of the detection limit (< 0.050 g/g), and this was obtained after one week. For other samples taken at the same time and in the same place but composed of the 0-5 cm fraction, residues over the detection limit were found up to the 12th week after treatment. The highest value detected was 0.301 g/g after 12 weeks. Samples taken after 10, 12 and 15 months showed no detectable residue. These results might be due to the chemical's poor vertical mobility in the soil. Altogether, 11 composite samples of raspberry and cherry foliage were collected over the first 8 weeks following treatment. The highest residue level was found in raspberry leaves, corresponding to 11.4 g/g fresh weight three weeks after treatment. After 8 weeks, the concentration had declined to only 1.01 g/g. At all times, lower concentrations were observed in cherry leaves than in raspberry leaves. For cherry twigs, the highest concentration detected among the 6 samples collected was 3.15 g/g fresh weight one week after treatment. After 12 weeks, the level was 2.25 g/g. Five berry samples were collected from raspberry bushes in all three sectors studied during the first two weeks after spraying. No more samples could be taken after that time since most of the fruit had fallen to the ground. The highest residue level detected was 28.7 g/g fresh weight after one week. The values observed in the second week were close to 17 g/g. By comparing the residues observed in this study with available toxicological data no toxic effect on exposed organisms are expected. However, residues in berries were higher than the maximum limit (0.1 ppm) established for a chemical in food according to the Canadian Food and Drugs Act, and for this reason, the Ministry is posting notices banning berry-picking at treated sites. 50. Lveill, P., J. Legris, and G. Couture. 1992. Rsidus de glyphosate dans les fruits sauvages la suite de pulvrisations terrestres en milieu forestier en 1989 et 1990. FQ92-3032. Government du Qubec, ministre des Forts, Service du suivi environmental. 25 p. In 1989 and 1990, the ministry of Forests continued the environmental monitoring of glyphosate (VisionTM) in forest. This herbicide is used for ground spraying required in tending softwood plantations in public forests. The main purpose of this study is to increase our knowledge of the behavior of glyphosate residues in raspberries and blueberries after a spraying of 1.5 kg a.i./ha. Samples were taken from two stations to determine the behavior of residues in raspberries during the first several days following spraying. A total of 16 composite samples was collected during the first week and fourteen days after treatment. The concentrations found in the raspberries varied from 3.54 to 27.1 g/g (fresh weight). During the first week, the average residues from the two stations increased up to 24.9 g/g, depending on the length of time after spraying. After fourteen days, these residues stood at 12.24 g/g. For the sake of comparison, eleven composite raspberry and six composite blueberry samples were taken punctually in 15 different sectors. The level of residues in raspberries varied from 1.79 to 20.7 g/g. The highest values were found during the period extending from the fifth to the tenth day after spraying. Before and after this period, residues were less than 10 g/g. As for blueberries, the highest value was 7.62 g/g, five days after spraying, and the lowest value 1.64 g/g, after 25 days. Before performing the laboratory analysis a fraction of each of the 17 samples mentioned was washed with tap water to determine the effect of washing the fruit. Subsequent analyses showed a significant drop in residues in some cases while, contrary to all expectations, we noticed an increase in other samples. However, washing the berries did not allow us to reduce the quantity of glyphosate residues below the 0.1 g/g limit in any of the 17 samples analyzed. We should keep in mind that this limit corresponds to the maximum quantity of chemical products tolerated in food. This standard was set by a regulation of Health and Welfare Canada. That is why the gathering of wild berries is prohibited until the end of the growing season in the areas sprayed. However, although the quantity of residues exceeds the standard, it is improbable that this amount has appreciable negative impacts on wildlife and humans. This is true even if we consider the highest level of residues found in the berries since the toxicity of glyphosate is very slight.

139

Plant and Soil Residues 51. . 1993. Reprsentativit d'un chantillon de sol la suite de pulvrisations ariennes et terrestres de glyphosate en milieu forestier. FQ93-3029. Gouvernement du Qubec, ministre des Forts, Direction de l'environnement, Service du suivi environnemental. 20 p. In 1988, the ministre des Forts continued its environmental follow-up program of glyphosate (VisionTM) in the forest environment. This herbicide is sprayed to maintain softwood plantations in public forests. The primary goal of this study was to verify whether or not a composite soil sample would reflect the level of glyphosate residues in a treated sector so that the product's possible effects in this environment could be evaluated. The study's second aim was to learn more about glyphosate persistence fifteen months after an application of 1.5 kg a.i./ha. Each sample analyzed was made up of a mix of ten soil cores taken from the 0-10 cm stratum. The sampling station was perpendicular to the axis followed by the spraying apparatus, and the cores collected were distributed equally in two treated strips. Two sectors treated by a ground sprayer and one sector sprayed from the air were visited. Five stations were set up randomly in each of these sectors. To compare the results, five other stations were set up along a transect in one of the sectors treated by a ground sprayer. A total of 35 soil samples were collected, one and six weeks after treatment. In one of the sectors treated by a ground sprayer, the five samples collected after one week contained an average of 0.358 g/g 0.268 (standard deviation) of glyphosate residues with a coefficient of variation (CV) of 79%. After six weeks, glyphosate residues were 0.359 g/g 0.293 (CV 86%). In this same sector, samples taken along the transect contained an average of 0.311 g/g 0.112 (CV 38%) and 0.234 g/g 0.075 (CV 34%) of residues after one and six weeks, respectively. In the second sector treated by a ground sprayer, on average, 1.40 g/g 0.744 (CV 56%) of residues were found in the five samples after one week. In the sector sprayed from the air, average residue content was 0.204 g/g 0.068 (CV 35%) after one week and 0.284 g/g 0.253 (CV 94%) after six weeks. Although the data show a degree of variability, our results allow us to conclude that this sampling method provides a rough estimate of the residues present in a given sector. This information is adequate for evaluating the effects of glyphosate residues in soil. Fifteen other samples were collected fifteen months after treatment to monitor the product's persistence. Six of them contained no detectable residues (<0.050 g/g), and the average for all the samples was 0.188 g/g. Levesque, C. A., J. E. Rahe, and D. M. Eaves. 1987. Effects of glyphosate on Fusarium spp.: its influence on root colonization of weeds, propagule density in the soil, and crop emergence. Canadian Journal of Microbiology 33: 354-60. Glyphosate is a broad spectrum herbicide that can lead to root rot like damage on crops. This study was undertaken to investigate the effect of glyphosate on the rootcolonizing Fusarium spp. The research was conducted at two sites. Site one was densely covered with perennial weeds, and site two with annuals. At site one, spraying the weed cover with glyphosate increased (p < 0.05) the level of colonization by Fusarium spp. in Ranunculus repens and Holcus lanatus, but not in Stellaria media and Plantago lanceolata. At site two, glyphosate enhanced colonization in Sperqula arvensis, Stellaria media, Echinochloa crusgalli, and Chenopodium albuim, but not in Capsella bursa-pastoris and Polygonum persicaria. At both sites, the number of colonyforming units of Fusarium spp. per gram of dried soil was increased by the application of glyphosate. Nevertheless, crops subsequently sown in the field containing the annual weeds were not detrimentally affected by glyphosate treatment of these weeds. . 1992. The effect of soil heat treatment and microflora on the efficacy of glyphosate in seedlings. Weed Research 32: 363-73. Seedlings of wheat (Triticum aestivum L.) and beans (Phaseolus vulgaris L.) were less sensitive to glyphosate when grown in heat-treated soil than in raw soil. Pythium spp. and Fusarium spp. were not detected in heat-treated loam or muck soils at the time of glyphosate treatment, although fungi of several other genera were present. The efficacy of glyphosate on wheat or beans grown in heat-treated loam soil was restored when untreated aqueous soil extracts were added to the heattreated soil. Bean seedlings grown in five different soil types varied in their sensitivity to glyphosate. 140

52.

53.

Plant and Soil Residues The variation in LD50 among autoclaved soils was lower than that among raw soils. Between 13- to 47-fold more glyphosate was required to kill the bean seedlings in any of the autoclaved soils compared with their corresponding raw soils. This differential effect was not observed on bean seedlings sprayed with either 2,4-D or paraquat. LD50 values for glyphosate on apple (Malus domestica Borkh.) seedlings growing in previously sterilized loam soil were reduced by inoculation of the soil with a representative Fusarium sp. or Pythium sp. obtained earlier from apple seedlings treated with glyphosate, but not by a Cylindrocarpon sp. from apple or Pythium ultimum Trow from glyphosatetreated bean. The efficacy of glyphosate on bean, wheat or apple seedlings can be affected by changes in certain microbial components of the soil. 54. Liu, L., Z. K. Punja, and J. E. Rahe. 1992. Phytoalexin production in bean roots grown on sterile media and in natural soil. Annual meeting of the American Phytopathological Society, Phytopathology , 82:1085. Application of the herbicide glyphosate at sublethal doses has been previously shown to enhance infection by Pythium and Fusarium spp. on bean roots, and reduced the levels of phaseollin, phaseollinisoflavan, phaseollidin and kievitone in bean leaves. The purpose of this research was to determine if glyphosate influences phytoalexin production in roots, which in turn could influence root susceptibility to fungal infection. Production of phytoalexins in healthy roots grown in sterile growth media including 0.65% water agar, silica sand and metro-mix was compared to that in natural soil. Seedlings were grown for 7-10 days, and the phytoalexins were extracted in 95% boiling ETOH, and quantitative analysis was performed by HPLC. Only a trace amount of phaseollin was found in roots grown on water agar. Large amounts of phaseollinisoflavan and phaseollin were produced in roots grown in metro-mix and natural soil. Both phytoalexins were also found in roots grown on silica sand, but at a lower concentration. The effects of glyphosate on altering the concentration or ratios of phytoalexins produced in bean roots are being investigated. Lonsjo, H., J. Stark, L. Torstensson, and B. Wessen. 1980. Glyphosate: decomposition and effects on biological processes in soil. Weeds and Weed Control, 21st Swedish Weed Conference, Vol. 1/2:140-146. (Sweden). Mamarbachi, G., N. Nadeau, and M. Carmichael. 1989. "Mthodes d'analyse du glyphosate dans l'eau, l'air, le sol, les sdiments, la vgetation et les tissues animaux. " Government of Quebec, Ministry of Energy and Resources, Environmental studies. 19 p. Marrs, R. H., A. J. Frost, and R. A. Plant. 1991. Effects of herbicide spray drift on selected species of nature conservation interest: The effects of plant age and surrounding vegetation structure. Environmental Pollution 69: 223-36. There has been increasing awareness of the possible impact of herbicide drift on vegetation in nature reserves and field-margin habitats adjacent to treated areas. However, relatively little is known about the impact of such drift on species typical of these habitats. To investigate this problem a series of bioassay experiments simulating spray drift were carried out with five native plant species (Digitalis purpurea, Leontodon hispidus, Lotus corniculatus, Lychnis flos cuculi, and Primula veris) of different age placed at different distances up to 4 m downwind from a sprayer under standardized conditions. These experiments used three herbicides.sbd.glyphosate, MCPA and mecoprop.sbd.in three types of surrounding vegetation structure.sbd.short, medium-height and tall grassland. Many plants showed symptoms of damage after spraying but showed no significant growth reduction at the end of the season even underneath the sprayer. Where a reduction in yield was found, it occurred close to the sprayer. In general, young plants were more often affected than old ones. Yield promotion occurred for some species between 2 and 4 m downwind of the sprayer (curvilinear response) with unknown ecological consequences. The structure of the surrounding vegetation influenced the response for some species, which indicates that deposition patterns can be complex, and thus there may be difficulty in predicting effects in semi-natural communities from simple deposition models. (United Kingdom).

55.

56.

57.

141

Plant and Soil Residues 58. Marsh, J. A. P., H. A. Davies, and E. Grossbard. 1977. The effect of herbicides on respiration and transformation of nitrogen in 2 soils. Part 1 metribuzin and glyphosate. Weed Research 17: 7782. The effects of metribuzin and glyphosate at 100 ppm on microbial C02 evolution and nitrogen transformation in 2 soils were investigated in the laboratory. Both herbicides reduced CO2 evolution from Boddington Barn soil (organic carbon content 1.5%, pH 6.6) at some dates, but neither gave any consistent effects on Triangle soil (organic carbon content 4.0%, pH 5.1). Both metribuzin and glyphosate stimulated mineralization of nitrogen for at least 9 weeks. Only metribuzin on Triangle soil gave any indication of inhibition of nitrification. Metribuzin degraded more rapidly in Triangle soil than in Boddington Barn. (United Kingdom). Martens, D. A., and J. M. Bremner. 1993. Influence of herbicides on transformations of urea nitrogen in soil. Journal of Environmental Science and Health Part B Pesticides Food Contaminants and Agricultural Wastes 28 , no. 4: 377-95. The influence of 28 formulated herbicides on transformations of urea nitrogen in soil was studied by determining the effects of 5 and 50 mg active ingredient/kg soil of each herbicide on the amounts of urea hydrolyzed and the amounts of nitrite and nitrate produced when samples of two coarse-textured soils and two fine-textured soils were incubated aerobically at 20 degrees C for various times after treatment with urea. The herbicides studied were alachlor, amitrole, atrazine, bentazon, bifenox, bromoxynil, chloramben, chlorpropham, 2,4-D amine, 2,4-D ester, dalapon, dicamba, dinitramine, dinoseb, diuron, EPTC, glyphosate, linuron, monuron, paraquat, pendimethalin, picloram, prometryn, propanil, propham, simazine, trifluralin and vernolate. When the 28 herbicides studied were applied at the rate of 5 mg active ingredient/kg soil, they did not retard urea hydrolysis in the four soils used or nitrification of urea nitrogen in the two fine-textured soils, but six of them (amitrole, chlorpropham, 2,4-D amine, dinoseb, propham and propanil) retarded nitrification of urea nitrogen in the two coarse-textured soils. When the herbicides were applied at the rate of 50 mg active ingredient/kg soil, two of them (alachlor and dinoseb) retarded urea hydrolysis in both of the coarsetextured soils, and six of them (bromoxynil, 2,4-D ester, dicamba, linuron, diuron and glyphosate) retarded urea hydrolysis in one of these soils. All of the herbicides tested retarded nitrification or urea nitrogen in one of the coarse-textured soils when they were applied at the rate of 50 mg active ingredient/kg soil, but only four of them (propanil, propham, amitrole and dinoseb) markedly retarded nitrification of urea nitrogen in all four of the soils used when they were applied at this rate. ANOVA and correlation analysis indicated that the effects of the herbicides tested on nitrification of urea nitrogen in soils were significantly influenced by incubation time, soil texture and soil organic-carbon content. The work reported indicates that the 28 herbicides studied are not likely to substantially affect nitrification when they are applied to soils in conjunction with urea fertilizers. Martensson, A. 1992. Assessing anthropogenic impact on heterotrophic nitrogen fixing soil microorganisms. Proceedings of the International Symposium on Environmental Aspects of Pesticide Microbiology J.P.E. Anderson et al. (editors). pp. 115-20. (Sweden). Martensson, A. M. 1992. Effects of agrochemicals and heavy metals on fast-growing Rhizobia and their symbiosis with small-seeded legumes. Soil Biology and Biochemistry 24: 435-45. The effect of potentially hazardous agrochemicals including fungicides, herbicides and heavy metals on symbiotic nitrogen fixation (in Trifolium pratense, Medicago sativa, Lotus corniculatus) have been investigated. The substances were tested with eight rhizobial strains from three cross-inoculation groups: Rhizobium leguminosarum b.v. trifolii, R. meliloti and R. loti in pure culture studies. Bacteria were obtained from a culture collection or from soils. Sensitivity of the bacteria to the agrochemicals and heavy metals varied. None of the bacteria were tolerant to all chemicals. No difference in tolerance between cross-inoculation groups existed. Bacteria were able to multiply at concentrations of agrochenmicals equal to or higher than recommended field-application rates. Heavy metals concentrations that severely inhibited growth were far lower than the highest amounts allowed under the current Commission of the European Communities' guidelines for environmental protection. Bacterial growth in presence of the agrochemicals and heavy metals, apart from glyphosate and zinc, did not influence nodulation ability of the strains. Development of uninoculated plants was inhibited at 142

59.

60.

61.

Plant and Soil Residues increasing concentrations of all compounds, red clover being most sensitive. Herbicides were most harmful, with injuries occurring at levels 1/10-1/10,000 of recommended applied concentrations. Uninoculated plants were less tolerant to agrochemicals, but were more tolerant to heavy metals compared to the bacteria. Root hair deformations similar to bacterial-induced root hair deformations were induced by bentazone, chlorsulphuron and monochlorophenoxyacetic acid on uninoculated plants. Symbiotic interactions were adversely affected by several of the agrochemicals. Bacterial-induced root hair deformations necessary for nodulation decreased in the presence of benomyl, bentazone, chlorsulphuron, fenpropimorph, mancozeb and monochlorophenoxyacetic acid. Fenpropimorph and mancozeb did not cause root hair deformations at increasing concentrations, indicating that these may inhibit nodulation under field conditions. Nodule development was inhibited at increased levels of bentazone, chlorsulphuron, glyphosate and mancozeb. Dry matter production of nodulated plants was adversely affected by bentazone and chlorsulphuron, indicating disturbances in nodule function. (Sweden). 62. Mattern, G. C., C. H. Liu, J. B. Louis, and J. D. Rosen. 1991. GC/MS and LC/MS determination of 20 pesticide residues for which dietary oncogenic risk has been estimated. Journal of Agricultural and Food Chemistry 39: 700-704. A rapid analytical procedure is described for the determination of 20 of 28 pesticides proposed by the National Research Council to pose a dietary oncogenic risk. Recovery and sensitivity studies were performed on a variety of food plants (including spinach, beans, peppers, lettuce and maize) for the suspected oncogens acephate, alachlor, azinphos-methyl, captafol, captan, chlordimeform, chlorothalonil, cypermethrin, diclofop-methyl, ethalfluralin, metolachlor, oxadiazon, parathion, permethrin, probanide, o-phenylphenol, terbutryne, folpet, linuron and oryzalin. All pesticides were determined using GC and chemical ionization mass spectrometry, except for the last 3, for which HPLC and MS were used. Average recoveries at the 0.5 p.p.m. fortification level were between 70 and 123%, with an average coefficient value of 13%. Sensitivity studies demonstrated that most pesticides could be detected at 0.05 and 0.10 p.p.m. in the crops, but some limits of detection were 0.25 p.p.m. or greater. Mekwatanakarn, P., and K. Sivasithamparam. 1987. Effect of certain herbicides on saprophytic survival and biological suppression of the take-all fungus. New Phycologist 106: 153-60. Saprophytic survival and pathogenicity of Gaeumannomyces graminis (Sacc.) ARX and Olivier var. tritici Walker were evaluated in sterile and unsterile western Australian wheat field soil treated with the herbicides glyphosate, Diquat + Paraquat, or Trifluralin. Survival in colonized straws and subsequent pathogenicity were not affected by the herbicide treatments in sterile soil. In unsterile soil, however, survival and pathogenicity were higher in glyphosate or Diquat + Paraquat-treated than in untreated or Trifluralin-treated soil. Incorporation of untreated natural soil into fumigated soil, to give concentrations of 1 and 10% natural soil in the mix, reduced disease in comparison with that in 100% sterile or 100% natural soil. Disease in fumigated soil increased progressively with increasing concentrations of glyphosate-treated, unsterile soil. Pre-treatment of wheat plants or soil with glyphosate before exposure of the host to the fungal pathogen or to other soil micro-organisms showed that the increase in disease following glyphosate treatment was not related to the direct effect of the herbicide on the host. (Australia). . 1987. Effect of certain herbicides on soil microbial populations and their influence on saprophytic growth in soil and pathogenicity of take-all fungus. Biology and Fertility of Soils 5: 175-80. The application of Diquat + Paraquat, glyphosate and Trifluralin to unsterilized field soil increased take-all caused by the fungus, Gaeumannomyces graminis var. tritici Walker by 13.0%, 16.6% and 10.8% respectively, while no effect on disease was recorded in sterilized soil treated with the same herbicides. The herbicides tested had no effect on the saprophytic growth of the pathogen with the exception of glyphosate, which increased its growth in unsterilized soil. The application of Diquat + Paraquat and glyphosate to unsterile soil had no effect on the numbers of Actinomycetes. The Diquat + Paraquat treatment, however, increased populations of fungi while the glyphosate decreased the numbers of bacteria. The proportion of soil fungi antagonistic to the pathogen was reduced in glyphosate-treated soil. The frequency of occurrence of Eupenicillium euglaucum (V.

63.

64.

143

Plant and Soil Residues beyma) Stolk and Samson (strain B), and Penicillium verruculosum Peyr. (strain B), which were strong and low level antagonists of GGT on agar, were reduced in their occurrence in soil by 7.7% and 2.5% respectively, following glyphosate treatment. Moreover, the numbers of Aspergillus viridi-nutans Ducker and Thrower, which showed moderate antagonism to the pathogen, was decreased by 1.9% and 4.1% in Diquat + Paraquat and glyphosate treatments respectively. The proportion of antagonists rather than total numbers of fungi appears to be related to the treatment effect observed on the soil growth and pathogenicity of G. graminis var. tritici in our investigation. The increase in disease of wheat in certain herbicide-treated soils may be due to the shift in soil microbial populations away from those which are antagonistic to the pathogen. (Australia). 65. Michailides, T. J., and R. A. Spotts. 1991. Effects of certain herbicides on the fate of sporangiospores of Mucor piriformis and conidia of Botrytis cinerea and Penicillium expansum. Pesticide Science 33: 11-22. The effects of several herbicides used in pome fruit orchards on the germination of spores and growth of mycelia of Botrytis cinerea, Mucor piriformis and Penicillium expansum in vitro and the survival of propagules of these fungi in soil were studied. Diuron in agar at 4-128 g/ml reduced germination of spores of B. cinerea and M. piriformis, and 2,4-D and paraquat at 32 g/ml similarly affected B. cinerea and P. expansum. Several herbicides at 128 g/ml in agar reduced growth of B. cinerea and M. piriformis but were ineffective against P. expansum. Propagule survival levels of the three fungi generally were lower in both autoclaved and non-autoclaved soil amended with herbicides than in non-amended soil. This effect was greatest in non-autoclaved soil, suggesting involvement of microbial antagonists. The most effective herbicides for reduction of fungal propagules in soil were 2,4-D, diuron, and paraquat. Miles, C. J., and H. A. Moye. 1988. Extraction of glyphosate herbicide from silt and clay minerals and determination of residues in soils. Journal of Agricultural and Food Chemistry 36: 486-91. Extraction of glyphosate from clay minerals and 2 soils was studied by a batch equilibrium technique. Experiments showed that for most of the sorbent/solvent systems studied, the amount of herbicide extracted increased as pH increased, suggesting that sorption occurs through ion exchange and hydrogen bonding. Alkaline solutions of kaolinite, iron oxide, and two Calvin silt loam soil samples showed resorption of glyphosate after an initial desorption step. Determination of glyphosate residues in soils was performed by derivatization of soil extracts with 9-fluorenylmethyl chloroformate and analysis by HPLC with fluorometric detection. Glyphosate was successfully recovered from sandy soils at 0.5 ppm by triplicate extraction with 0.1 M KH2PO4 and 30-fold concentration. For soils with high clay content, acceptable recovery at 1.0 ppm was achieved with triplicate extraction using 0.2 M KOH. Miller, J. J., N. Foroud, B. D. Hill, and C. W. Lindwall. 1992. Herbicides in surface runoff soil and ground-water under different agricultural management practices in southern Alberta. Canadian Journal Soil Science 72: 329. Miller, J. J., B. D. Hill, C. Chang, and C. W. Lindwall. 1995. Residue detections in soil and shallow groundwater after long-term herbicide applications in southern Alberta. Canadian Journal of Soil Science 75, no. 3: 349-56. After herbicide applications for 1-24 yr, there were no detectable residues of glyphosate, dicamba, 2,4-D, bromoxynil or methylchlorophenoxyacetic acid (MCPA) in soil at two long-term tillage sites and one long-term manured site. The only detectable residues in soil were of diclofop and triallate. Residues of bromoxynil, diclofop and MCPA but not dicamba, 2,4-D or triallate, were detected in the groundwater at the manured site. Diclofop was detected in 6% and bromoxynil and MCPA in 2% of 84 water samples collected at the manured site. Maximum concentrations of bromoxynil (6.5 mu g L super (-1)) and diclofop (47 mu g L super (-1)) in the groundwater at the manured site exceeded levels set by the Canadian drinking water guidelines. Long-term application of herbicides has not caused accumulation of harmful residues in southern Alberta soils, but the presence of certain herbicides in the groundwater at concentrations above the level set by the drinking water guidelines is cause for concern.

66.

67.

68.

144

Plant and Soil Residues 69. Morash, R., and B. Freedman. 1989. The effects of several herbicides on the germination of seeds in the forest floor. Canadian Journal of Forest Research 19: 347-50. In a laboratory experiment, we studied the effects of perturbation of a forest floor substrate with six concentrations (10, 50, 100, 500, 1000 and 5000 ppm) of four herbicides: glyphosate (Nphosphonomethyl glycine), 2,4,5-T (2,4,5-trichlorophenoxy acetic acid), triclopyr (3,5,6-trichloropyridinyloxyacetic acid), and two formulations of hexazinone (3-cyclohexyl-6-(dimethylamino)-1-methyl1-1,3,5-triazine-2,4(1H,3H)dione). Although their toxic thresholds differ, the herbicides all caused significant decreases in the germination of seeds. However, large decreases in germination only occurred at concentrations that are unrealistically large in comparison with the herbicide residues that actually occur after a silvicultural treatment. In a parallel field experiment, no significant difference in seeding germination was observed for forest floor samples that were exposed to or shielded from herbicide deposition at two sites that were sprayed in an operational program. Morillo, E., C. Maqueda, M. Bejarano, L. Madrid, and T. Undabeytia. 1994. Cu (II)-glyphosate system: A study by anodic stripping voltammetry and the influence on Cu adsorption by montmorillonite. Chemosphere 28, no. 12: 2185-96. The influence of the pesticide glyphosate (GPS) on the adsorption of Cu(II) on montmorillonite has been examined. The complexation of Cu(II) with GPS was studied using anodic stripping voltammetry in differential pulse mode (DPASV). It has been concluded that the complexes present a labile behaviour and GPS shows a low but noticeable degree of heterogeneity, probably due to complexation of Cu by more than one GPS species. Cu(II) adsorption on montmorillonite is drastically decreased in the presence of GPS, due to several reasons: decrease in free Cu concentration due to formation of Cu-GPS complexes; surface loading of GPS on montmorillonite, obstructing interlamellar Ca super (2+) adsorption and competitive effect between protons and Ca super (2+) for interlamellar positions. (Spain). Muller, M. M., C. Rosenberg, H. Siltanen, and T. Wartiovaara. 1981. Fate of glyphosate and its influence on nitrogen cycling in two Finnish agriculture soils. Bulletin of Environmental Contamination and Toxicology 27: 724-30. (*) During an energy forest experiment in Suomusjarvi, Kettula (located in Southern Finland) glyphosate was used to destroy indigenous flora, mainly quackgrass, before replanting. Since the degradation rate of glyphosate and its biological impact are known to vary and field data were largely unavailable, it was decided that both the persistence of glyphosate and its effects on the nitrogen cycling processes fundamental to the energy forest experiment should be investigated. Both fields, SK1 (loam soil) and SK2 (fine silt soil), were treated on the 14th of September, 1978 with glyphosate 2.6 kg/ha. Two sites of ca. 100 m2 were chosen at each field and sampled immediately after the application of glyphosate, 1 month and eight months later. Control samples were collected before glyphosate was applied. It is concluded that glyphosate degraded even at low temperature conditions, and it is not to be expected that the application of glyphosate will affect directly nitrogen fixation, nitrification or denitrification activity in these soils. (Finland). Murray, B., and J. Taylor. 1991. The Codex Committee on pesticide residues. Pesticide News 11: 12-15. The primary objectives of the Codex Committee, one of the standards organizations sponsored by the United Nations WHO and FAO, are described, namely the establishment of maximum residue limits (MRLs) that can be accepted by other countries for pesticides in food and animal feeds in order to facilitate trade and protect the health of consumers. Highlights of the Apr. 1990 meeting of the Committee are presented, including discussion of the MRLs of endosulfan, cyhexatin, dinocap, daminozide, dithiocarbamates, aldicarb, amitraz, procymidone, glyphosate and terbufos, of general residue limits for fruit and vegetables and of fumigants. Issues arising that are important in the setting of worldwide standards for pesticide residues in food and other commodities are also discussed.

70.

71.

72.

145

Plant and Soil Residues 73. Newton, M., L. M. Horner, J. E. Cowell, D. E. White, and E. C. Cole. 1994. Dissipation of glyphosate and aminomethylphosphonic acid in North American forests. Journal of Agricultural and Food Chemistry 42, no. 8: 1795-802. Residues of glyphosate (N-phosphonomethylglycine) and its metabolite aminomethylphosphonic acid (AMPA) were followed on three forested sites in Oregon, Michigan, and Georgia. Eight-hectare residual stands of low-quality hardwoods were treated with 4.12 kg/ha glyphosate a.e. applied aerially in late summer. Residues were highest in upper crown foliage. Overstory reduced exposure of understory vegetation and streams. Residues in streams were close to the detection limit or undetectable in 3-14 days. Residues in soils were highest where cover was sparse and where litter was removed. No residues were detectable in soil 409 days after treatment; movement below 15 cm was negligible. AMPA appeared at low levels in all degrading matrices, including sediments soon after deposition of glyphosate. In pond sediments, both glyphosate and AMPA remained bound and inactive. Residue concentrations in foliage, water, and soil were below levels known to be biologically active in nontarget fauna. Newton, M., K. M. Howard, B. R. Kelpsas, R. Danhaus, C. M. Lottman, and S. Dubelman. 1984. Fate of glyphosate in an Oregon USA forest ecosystem. Journal of Agricultural and Food Chemistry 32: 1144-51. See Mammals Section. Ogner, G. 1987. Glyphosate application in forest-ecological aspects. II. The quality of water leached from forest soil lysimeters. Scandinavian Journal of Forest Research 2: 469-80. See Water Quality Section. . 1987. Glyphosate application in forest-ecological aspects. III. The chemical characterization of four Norwegian forest soils. Scandinavian Journal Forest Research 2: 48198. Chemical parameters of soil profiles of four different forest soils were investigated after glyphosate application in a lysimeter experiment lasting 28 months. Generally, there was hardly any effect found on the nutrient status of the soils due to glyphosate treatment. A slight increase in NO3and a decrease in pH and extractable Na+ were detected. The amounts of extractable NH4+ and org.-N, t-N, extractable K+, Mg2+, Ca2+, Mn2+, Fen+and Aln+, CEC and BS in the soil profiles were not affected by glyphosate application. These results, taking into consideration the extreme treatment of the soil, justify the conclusion that no harmful ecological effects on forest soil chemical status due to glyphosate application are likely. (Norway). Olson, B. M., and C. W. Lindwall. 1991. Soil microbial activity under chemical fallow conditions effects of 2,4-D and glyphosate. Soil Biology and Biochemistry 23: 1071-76. Field and laboratory studies were made to examine the effects of 2,4-D and glyphosate on soil microbial activity under zero-tillage chemical fallow conditions. Glyphosate and 2,4-D applied three consecutive times at field rates and 10 times the field rates had no effect on microbial biomass-C, C mineralization or nitrification in the field. Laboratory studies showed that 2 and 100 times the field rate of 2,4-D reduced nitrification by 11 and 79%, respectively. In laboratory studies, 2,4-D incorporated into the soil reduced nitrification, but surface-applied 2,4-D did not. Glyphosate and 2,4-D used as recommended should not affect microbial activity under dryland zero-tillage chemical fallow conditions. Pawlizki, K. H. 1987. Effects of herbicides used in orchards and vineyards on man and the environment. Gesude Pflanzen 39: 486-96. Residual effects of herbicides used in orchards and vineyards are reviewed in relation to the influence of soil sorption, clay and OM content, soil moisture, climatic factors, repeated herbicide treatment, interactions between different herbicides, persistency, and accumulation in the soil, residue binding to soil components and toxicological effects. These include contamination of the groundwater and side-effects on the soil fauna and flora, insects, fishes and man. It was concluded that there is currently no indication of damage to the chemical and biological processes in the soil in orchards and vineyards. Also, herbicide residues in the fruit were not considered a threat to health, but the risk of exposure to herbicides during spraying operations was emphasized. Groundwater contamination by 146

74.

75.

76.

77.

78.

Plant and Soil Residues paraquat, diquat and glyphosate was ruled out due to a long time interval between application and harvesting, but contamination by simazine and Caragard 3587 (terbuthylazine + terbumeton) was considered possible under unfavourable soil and climatic conditions. (Germany). 79. Phillips, S. F., D. L. Wotton, D. Jones, and D. B. McEachern. 1987. Glyphosate residues in vegetation and soils of Manfor sites after Roundup application in August, 1985. Manitoba Department of Environment and Workplace Safety and Health, Environmental Management Services Branch, Terrestrial Standards and Studies. Report No. 87-2. 30 p. In August 1985 Manfor applied Roundup herbicide aerially (helicopter) to three forest blocks in the area of The Pas, Manitoba. The objective was to test its effectiveness in controlling hardwood competition in softwood plantations and also to test its effectiveness as a site preparation tool. Application rates of 3, 4 and 5 l/ha were tested. Terrestrial Standards and Studies of the Department of Environment and Workplace Safety and Health, Province of Manitoba, conducted monitoring to determine the distribution, fate and persistence of the active ingredient, glyphosate, in the environment. The amount of glyphosate intercepted by mylar plates was used to estimate actual spray deposition. Same day, 32 day and 295 day post-spray samples of aspen foliage and soil were analyzed for glyphosate residues. The 3 l/ha rate was not as effective in killing vegetation as the 4 or 5 l/ha rate. The mylar deposition monitors provided good estimates of actual spray distribution. Initial concentrations of glyphosate in aspen leaves should not be toxic to birds or animals which might ingest them. Glyphosate residues deposited on the soil were adsorbed entirely by the surface organic material and no downward leaching occurred. Rapid breakdown of soil residues occurred. The half-life of glyphosate in forest soils of The Pas area appears to be less than 16 days. Piccolo, A., and G. Celano. 1994. Hydrogen-bonding interactions between the herbicide glyphosate and water-soluble humic substances. Environmental Toxicology and Chemistry 13, no. 11: 1737-41. See Water Quality Section. Piccolo, A., G. Celano, M. Arienzo, and A. Mirabella. 1994. Adsorption and desorption of glyphosate in some European soils. Journal of Environmental Science and Health Part B Pesticides Food Contaminants and Agricultural Wastes 29, no. 6: 1105-15. The interaction of glyphosate (N- (phosphonomethyl) -glycine) with four typical European soils is reported. Results of adsorption and desorption isotherms show that the interaction of glyphosate with these soils was mainly related to content of iron and aluminium amorphous hydroxides. Moreover, it was found that the presence of divalent cations in 2:1 clay minerals also contribute to glyphosate adsorption. The S-type form of the adsorption isotherms revealed the existence of two different binding sites. These were exchangeable cations at low herbicide concentration and Fe and Al at higher glyphosate concentrations. The K maximum values of adsorption provided by the linear form of the Langmuir equation were found to be more consistent with soil parameters than those calculated by the Freundlich equation. The order of desorption from the soils was the reverse of that found for adsorption. Moreover, desorption varied from around 15 to 80% of the adsorbed herbicide according to the soil characteristics. This indicated that glyphosate adsorption on soils is far from being permanent and leaching to lower soil horizons with limited biological activity may occur. (Italy). Piccolo, A., L. Gatta, and L. Campanella. 1995. Interactions of glyphosate herbicide with a humic acid and its iron complex. Annali Di Chimica 85, no. 1-2: 31-40. (Italy). Preston, C. M., and J. A. Trofymow. 1989. Effects of glyphosate (Roundup) on biological activity of two forest soils. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 122-40. The effects of glyphosate (ROUNDUP) on some biological processes were examined in laboratory and field trials. In the laboratory study, forest floor material and mineral soil (0-5 cm depth) from a Douglas-fir site at Shawnigan Lake (Vancouver Island) were incubated with ROUNDUP (10 and 50 g a.i./g), with or without N15labelled urea (at 200 g/g N). No significant effects of ROUNDUP at either level of application were found on carbon dioxide evolution, urea hydrolysis, nitrification, or 147

80.

81.

82.

83.

Plant and Soil Residues immobilization of ammonium, for either forest floor or mineral soil layers. In field studies, soil fauna and microflora populations were monitored in surface organic layers prior to and following ROUNDUP application on alder-covered sites at Carnation Creek (Vancouver Island). In a 6-month study, ROUNDUP had no significant longterm effects on populations of soil fauna or microflora. In an intensive 1-month field sampling trial, soil microflora populations in ROUNDUP-treated plots fluctuated and then returned to control levels. 84. Putnam, A. R. 1976. Fate of glyphosate in deciduous fruit trees. Weed Science 24: 425-30. Glyphosate [N-(phosphonomethyl)glycine] was applied to selected areas on young apple (Malus sylvestris L.), pear (Pyrus communis L.), sour cherry (Prunus cerasus L.) and peach [Prunus persica (L.) Patsch] trees to observe tree response. Injury occurred only on newly planted peach when contact was made on the basal 12.7 cm of tree trunks. Glyphosate effectively killed suckers without damaging trees. When sprays were applied to a lower branch of apple trees, local injury occurred the same season and the following year, but symptoms were not visible in other portions of the tree. Radiolabeled glyphosate was applied to the basal trunk, leaves, suckers, and fruit of 4-year-old 'MacSpur'/MM106 apple and 5-year-old 'Bartlett'/seedling pear. Applications on basal trunk areas produced no detectable radioactivity in leaves, buds, or fruits at harvest time. Applications on sucker leaves produced radioactivity only in the treated and adjacent sucker tissue and not in other portions of the tree. The 14C-glyphosate moved readily from the treated leaves of lower branches to other leaves, buds, and developing fruit on the same branch but was not detectable in other areas of the tree. After 90 days, 92 to 98% of the extractable radioactivity in leaf, bud, or fruit tissue was unaltered 14Cglyphosate. Quinn, J. P., J. M. M. Pedden, and R. E. Dick. 1988. Glyphosate tolerance and utilization by the microflora of soils treated with the herbicide. Applied Microbiology and Biotechnology 29: 511-16. The effect of recurrent applications of the herbicide glyphosate on a garden soil was investigated. Compared to an adjacent untreated soil the microbial population showed reduced sensitivity to glyphosate when grown in mineral salts medium. In both populations inhibition could be partially reversed by addition to the medium of the end products of the aromatic amino acid biosynthetic pathway, but the effect was more pronounced in the population from the treated site. However, all isolates from both soils were capable of growth in unsupplemented medium in the presence of as much as 10 mM glyphosate. No evidence for glyphosate metabolism was obtained from enrichment experiments carried out using inocular from the untreated soil; at the treated site organisms capable of using glyphosate as sole C or N source could not be isolated but a variety of gram-negative bacteria able to use its phosphonate moiety were obtained. Many of these organisms were identified as Pseudomonas spp. (N. Ireland). Raatikainen, M., H. Siltanen, C. Rosenberg, T. Raatikainen, and J. Mukula. 1979. Herbicide residues on cowberries Vaccinium vitis idaea bilberries Vaccinium myrtillus and lichens in controlled ground spraying experiments on woodland. Annales Agricultureae Fenniae 18: 11216. (Finland). Reinert, K. H. 1989. Environmental behaviour of aquatic herbicides in sediments. Symposium on reactions and movement of organic chemicals in soils held at the 1987 annual meeting of the American Society of Agronomy and the Soil Science Society of America. SSSA Special Publication (Soil Science Society of America ) 22:335-48. Reynolds, P. E., J. C. Scrivener, L. B. Holtby, and P. D. Kingsbury. 1989. A summary of Carnation Creek herbicide study results. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 322-34. The Carnation Creek watershed, located on the west coast of Vancouver Island, was aerially treated with Roundup (glyphosate) from September 6 to 15, 1984 with the intent of examining environmental fate and impact of Roundup treatment on a temperate coastal rain forest. Carnation

85.

86.

87.

88.

148

Plant and Soil Residues Creek presents a unique research situation in that a 20 yr data base on its salmonid population exists. This data base has permitted whole life history impact assessment of the effects of numerous forest management practices, including the possible impacts of herbicide use on this resource. Following Roundup treatment at Carnation Creek, various chemical and biological studies were conducted for up to three years post-treatment. These studies revealed no unexpected or long-term adverse effects on coho salmon or other aquatic organisms using tributaries that had been directly over-sprayed with the herbicide. Residue movements within the watershed and residue inputs into the aquatic ecosystem were carefully monitored in relation to autumn and winter storms. Glyphosate residues rapidly dissipated and degraded in the natural environment. After one year remaining residues were strongly absorbed to organic matter, soil particles, and/or stream bottom sediments, were deemed to be biologically unavailable. 89. Rhodes, A. N., and C. W. Hendricks. 1990. A continuous-flow method for measuring effects of chemicals on soil nitrification. Toxicity Assessment 5: 77-90. Because the productivity of terrestrial ecosystems is directly related to microbial nutrient cycling, understanding the effects of chemical contaminants on soil microbial processes is important. This study examined the effects of two model chemicals-Roundup (glyphosate) and N-Serve (nitrapyrin)-on nitrifying organisms in static, perfusion, and continuous-flow culture systems. Experimental concentrations were approximately 1, 10, and 100 times the spot application rate. Both N-Serve and Roundup were shown to inhibit nitrification in the treated soils. Roundup significantly reduced nitrification at 6.8 and 68 mg/g dry soil. N-Serve (nitrapyrin) completely inhibited nitrification at levels greater than 42 m/g dry soil in all cultural methods. In comparative studies with static batch and perfusion culture techniques, the continuous-flow system proved to be both reliable and useful in the culture of nitrifying bacteria. This method provides an alternative to traditional culture techniques in measuring chemical effects on microbial geochemical cycles and provides a new method for use in toxicity testing. Roberts, W., and B. Cartwright. 1990. Alternative approaches to soil erosion, fertility and pest management of cabbage. 50th Annual Meeting of the American Society for Horticultural Science (Southern Region). Hortscience 25:859 Roslycky, E. B. 1982. Glyphosate and the response of the soil microbiota. Soil Biology and Biochemistry 14: 87-92. Limited information on the effect of glyphosate (N-phosphonomethylglycine) on soil microorganisms justified an inquiry into the response of soil actinomycetes, bacteria and fungi in terms of their respiration, and sensitivity of isolates. Low concentrations of glyphosate had little effect on total populations of these organisms during the 214-day experiment, while high concentrations initially increased actinomycete and bacterial numbers by 2 and 1 logs, respectively. The stimulation was followed by a decline and fluctuation showing a gradual increase in numbers. The respiration rates of the soil microbiota in soil suspensions, showed some irregular stimulation and retardation with up to 10 m.g glyphosate/ml. In contrast high doses suppressed 02 uptake by the microbiota. Fungi were the least affected. Pronounced inhibition of actinomycete and bacterial respiration was in agreement with the results from isolate replication. The results indicated both stimulation and inhibition of 02 uptake by some organisms within these groups. In contrast to some reports of limited, shortterm inquiries these results showed considerable effects of glyphosate on soil microorganisms. Roy, D. N., and S. K. Konar. 1989. Development of an analytical method for the determination of glyphosate and aminomethylphosphonic acid residues in soils by nitrogen-selective gas chromatography. Journal of Agricultural and Food Chemistry 37: 441-43. An analytical method has been developed for the determination of glyphosate and its metabolite (aminomethyl)phosphonic acid residues in soils. The herbicide and its metabolite were extracted with deionized water in the presence of phosphoric acid. The compounds in the extract were derivatized with a mixture of trifluoroacetic anhydride and trifluoroethanol and quantified by gas chromatography using a nitrogen-phosphorus detector. The limits of detection of glyphosate and its metabolite were 0.05 and 0.01 m.g/g.

90.

91.

92.

149

Plant and Soil Residues 93. Roy, D. N., S. K. Konar, S. Banerjee, D. A. Charles, D. G. Thompson, and R. Prasad. 1989. Persistence movement and degradation of glyphosate in selected Canadian boreal forest soils. Journal of Agricultural and Food Chemistry 37: 437-40. Persistence, mobility, and degradation studies of glyphosate, N-(phosphonomethyl)glycine, under actual field conditions of boreal forest soils of Ontario, were undertaken after spraying Roundup at the rate of 2 kg a.i./ha. Soils at three depths were collected and analyzed for residues of glyphosate and its metabolite (aminomethyl)phosphonic acid. Glyphosate was found to remain consistently to a level below 50% of the highest residue values observed beyond 24 days. More than 95% of the total herbicide residue was found in the upper organic layer at any time. There was no evidence of lateral movement of the glyphosate either in runoff water or through subsurface flow. In general, concentrations of the metabolite (aminomethyl)phosphonic acid were very low. . 1989. Uptake and persistence of the herbicide glyphosate (VisionR) in fruit of wild blueberry and red raspberry. Canadian Journal of Forest Research 19: 842-47. Field studies on the uptake and persistence of glyphosate (N-(phosphonomethyl)glycine) on wild blueberry (Vaccinium myrtilloides Michx.) and red raspberry (Rubus strigosus Michx.) under boreal forest (Matheson, Ont.) conditions were undertaken. Uptake studies indicated that less than 10% of glyphosate penetrated the fruit in the first 9 h postapplication. Results of the persistence studies showed a gradual decline in residue levels with time. Times to 50% dissipation for glyphosate residue as determined by curvilinear regression analyses were <20 days (95% confidence limit of 8-26 days) and <13 days (95% confidence limit of 6-14 days) for blueberry and raspberry fruit, respectively. Initial residue levels dissipated to approximately 4 and 6% after 61 and 33 days for the blueberry and raspberry, respectively. Results also showed that at no time during the study period did glyphosate levels in either substrate dissipate to below the maximum permissible residue level (0.01 ppm) as established by Health and Welfare Canada. Rueppel, M. L., B. B. Brightwell, M. Schaefer, and J. T. Marvel. 1977. Metabolism and degradation of glyphosate in soil and water. Journal of Agricultural and Food Chemistry 25: 517-28. Complete and rapid degradation of glyphosate (N-phosphonomethyl glycine (1)) occurs in soil and/or water microbiologically and not by chemical action. Using soil/water shake flasks, up to 50% of each carbon of 1-C14 was evolved as C14O2 in 28 days. In two of the three soils examined, 1 was 90% dissipated in less than 12 weeks. Aminomethyl phosphonic acid (2), the only significant soil metabolite of 1, also undergoes rapid degradation in soil. Short-term shake flask metabolism experiments with both C13- and C14-labelled 1 were carried out in order to permit facile, unequivocal spectral identification of 1 and its transient metabolite aminomethyl phosphonic acid (2). Comparison of the metabolic samples to both reference standards and the spiked controls by means of H1, P31, and C13 NMR, mass spectral analysis, ion-exchange chromatography, and thin-layer chromatography has unequivocally characterized both bound and unbound 1 and 2 in soil. The parent herbicide 1 has also been shown to be stable to sunlight, nonleachable in soil, to have a low propensity for runoff, and to have a minimal effect on microflora. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. Wageningen, NL: European Weed Res. Soc. pp. 315-21. See Water Quality Section. Sahid, I., M. A. Salleh, and O. Omar. 1993. Effects of herbicides on cellulose decomposition in soil. Plant Protection Quarterly 8, no. 2: 38-39. (Malaysia). Samuel, O., L. A. Ferron, and L. St-Laurent. 1996. valuation de lexposition cutane et estimation dun coefficient de transfert des rsidus foliaries dlogeables pour la population expose au glyphosate (english title: Evaluation of dermal exposure of a population exposed to glyphosate, and determination of a transfer coefficient for dislodgeable foliar residues.),

94.

95.

96.

97.

98.

150

Plant and Soil Residues Ministre des Ressources naturaelles du Qubec, Direction de lenvironnement forestier , Sainte-Foy, Centre de toxicologie du Qubec. 56 p. See Human Health Section. 99. Sanchez-Martin, M. J., T. Crisanto, M. Arienzo, and M. Sanchez-Camazano. 1994. Evaluation of the mobility of C-14-labelled pesticides in soils by thin layer chromatography using a linear analyzer. Journal of Environmental Science and Health Part B Pesticides Food Contaminants and Agricultural Wastes 29, no. 3: 473-84. The mobility of seven pesticides in a chromic cambisol soil was studied by soil thin layer chromatography. Pesticide mobilities were determined by means of conventional autoradiographs of the chromatograms, as well as from sequential series of curves and images of the pesticide spots provided by a linear analyzer. The R-f values obtained from the autoradiographs and those provided by the linear analyzer were quite consistent. Based on such values, pesticide mobility decreased in the following order: acephate gt fluometuron gt atrazine gt ethofumesate gt metolachlor gt diazinon gt glyphosate. According to the mobility scale proposed by Helling and Turner (1968), acephate is highly mobile; atrazine, fluometuron, ethofumesate and metolachlor are moderately mobile; diazinon is slightly mobile; and glyphosate is immobile. The images provided by the linear analyzer allow to determine the R-f values for the zones of maximum activity in the pesticide spots (R-f-max), as well as the activities of different spot zones and those corresponding to R-f and R-f-max. The results obtained show the image analyzer to provide more expeditious R-f measurements from the chromatograms and open up new prospects for using soil TLC to investigate pesticide mobility. Schuster, E., and D. Schroder. 1990. Side-effects of sequentially-applied pesticides on nontarget soil microorganisms: field experiments. Soil Biology and Biochemistry 22: 367-73. In trials in 1985-86, the effect of a plant protection system consisting of 7 pesticide treatments was measured on the microbial flora of a Gleyic Luvisol soil. Successive applications of 3 kg dinosebacetate (Aretit)/ha, 2.5 kg isoproturon (Arelon), 4 kg dichlorprop (U-46 DP), 0.5 kg chlormequat (Cyclocel), 1.5 kg prochloraz + carbendazim (Sportak Alpha), 4 kg captafol + triadimefon (Bayleton DF), 0.3 kg pirimicarb (Pirimor) and 4 kg glyphosate (Roundup) caused only slight and short-lived sideeffects on microbial activity which usually disappeared before the next treatment was applied. The impact of the pesticides on soil microflora appeared to be significantly affected by climatic conditions. (Germany). Shipman, R. D., and T. J. Prunty. 1988. Effects of herbicide residues on germination and early survival of red oak acorns. Proceedings, 42nd annual meeting of the Northeastern Weed Science Society , 86-91. The response of greenhouse-grown Quercus rubra to several herbicides at 1, 2 and 4 times the manufacturer's recommended rate was studied under greenhouse conditions. Dicamba, 2,4-D, picloram, sulfometuron-methyl and triclopyr plus or minus picloram significantly reduced the percentage of germinating acorns. Picloram, sulfometuron-methyl and triclopyr significantly reduced seedling height. Glyphosate plus or minus non-ionic wetter, hexazinone and simazine had little or no effect on percentage germination or seedling height. However, hexazinone inhibited photosynthesis and eventually killed the seedlings. Soil type significantly affected herbicide availability and subsequent phytotoxicity. Shuma, J. M., W. A. Quick, M. V. S. Raju, and A. I. Hsiao. 1995. Germination of seeds from plants of Avena fatua L. treated with glyphosate. Weed Research 35, no. 4: 249-55. Glyphosate was applied at four rates under greenhouse conditions to Avena fatua L. plants at four stages of seed development. Application at anthesis completely prevented the formation of viable seeds. Application five days after anthesis (DAA) of the terminal floret of the panicle significantly reduced seed production at all herbicide rates used, and at 1.76 kg a.i./ha no viable seeds were produced. When applied 10 DAA, only the highest rate of glyphosate resulted in substantial reduction in number of primary seeds, but seed viability suffered at all herbicide levels. Glyphosate applied 15 DAA still produced a significant decrease in primary and secondary seed production and biomass. Both the viability and the germination rate of seeds from treated plants were significantly affected. When the herbicide was applied to plants 5 DAA, no viable seeds were produced by plants surviving 151

100.

101.

102.

Plant and Soil Residues the highest rate, and all rates significantly reduced germination. Glyphosate applied 10 DAA significantly suppressed germination, with 1.76 kg a.i./ha being the most effective rate. When applied to plants 15 DAA, only the highest rate of glyphosate significantly affected the overall germination of both primary and secondary seeds, but the normal imposition of dormancy was partially blocked in seeds from plants treated with 0.44 and 0.88 kg a.i./ha. These findings are relevant to chemical summerfallow and crop desiccation practices. 103. Siltanen, H., C. Rosenberg, M. Raatikainen, and T. Raatikainen. 1981. Triclopyr, glyphosate and phenoxyherbicide residues in cowberries, bilberries and lichen. Bulletin of Environmental Contamination and Toxicology 27: 731-37. (*) The results from this and previous studies show that the residues in berries were of the same order of magnitude when the same amount of any of the studied herbicides was used. In aerial spraying, the application rates for the different herbicides vary; the lowest is for glyphosate, 0.5-0.75 kg/ha, and the highest for phenoxyherbicides, 2.5 kg/ha. The application rate for triclopyr is about 1.2 kg/ha. As a consequence of this the residues in berries from aerially sprayed forest can be expected to be highest after phenoxyherbicide application and lowest after glyphosate application. (Finland). Simpson, J. A., A. M. Gordon, P. E. Reynolds, R. A. Lautenschlager, F. W. Bell, D Gresch, and D. Buckley. 1997. Influence of alternative conifer release treatments on soil nutrient movement. Forestry Chronicle 73, no. (In press). Zero tension soil solution samplers were used to collect soil solution at 75 cm for one and onehalf years after the following treatments: 1) helicopter applied Release silvicultural herbicide (a.e. triclopyr @ 1.9 kg a.e. ha-1), and 2) helicopter applied Vision silvicultural herbicide (a.e. glyphosate @ 1.5 kg a.e. ha-1); 3) motor-manual cutting (brushsaws); and 4) control (no treatment). Results show no substantial treatment related differences in the movement of selected nutrients (total organic N, NH4+, NO3-, K, Ca) among these treatments. Mean nitrate concentration was less than 1.5 ug mL-1 throughout the sampling period and was marginally higher, but not statistically different, following release treatments. Smith, A. E., and A. J. Aubin. 1993. Degradation of super(14)C-glyphosate in Saskatchewan soils. Bulletin of Environmental Contamination and Toxicology 50, no. 4: 499-505. The following study was undertaken to investigate the breakdown of super (14)C-glyphosate in three Saskatchewan soils, under laboratory conditions using a closed system, by measuring amounts of super (14)CO2 evolved over a 90-day period. After 90 days, solvent extractable radioactivity and nonextractable super(14)C associated with the soil was also quantified. Smith, E. A., and F. W. Oehme. 1992. The biological activity of glyphosate to plants and animals: A literature review. Veterinary and Human Toxicology 34 , no. 6 : 531-43 . This report reviewed the properties of glyphosate (1071836), which is the active ingredient of several herbicides. Specific topic covered in this review included: the physical and chemical properties of glyphosate; use of the herbicide; glyphosate storage and removal or disposal; mobility and inactivation in soil; mechanisms of inactivation in soil; phosphate reversal of glyphosate binding; microbial, chemical and photolytic degradation; effects on soil enzymes; leaching in soil; absorption into plants; translocation in plants; mode of herbicidal action; toxic effects on plants; mutagenic effects; teratogenic and reproductive effects; absorption and excretion; biotransformation; distribution; biochemical studies on metabolites; effects in animals; phototoxicity; skin sensitization; skin and eye irritation; short and analytical detection; and treatment of exposures. The authors conclude that when used correctly, glyphosate poses no threat to the environment and its inhabitants.

104.

105.

106.

152

Plant and Soil Residues 107. Soulas, G. 1992. Biological availability of pesticides in soil 2,4-D and glyphosate as test cases. Proceedings of the International Symposium on Environmental Aspects of Pesticide Microbiology J.P.E. Anderson et al. (editors). pp. 219-24. (Sweden). Sprankle, P., W. F. Meggitt, and D. Penner. 1975. Adsorption, mobility and microbial degradation of glyphosate in the soil. Weed Science 23: 229-34. Glyphosate [N-(phosphonomethyl)glycine] was readily bound to kaolinite, illite, and bentonite clay and to charcoal and muck but not to ethyl cellulose. Fe+3 and Al+3-saturated clays and organic matter adsorbed more glyphosate than Na+ or Ca+-saturated clays and organic matter. Glyphosate appears to be bound to the soil through the phosphonic acid moiety as phosphate in the soil competed with 14C-glyphosate for adsorption sites. Glyphosate mobility in the soil was very limited and was affected by pH, phosphate level, and soil type. The 14C-glyphosate was biodegraded in soil to 14CO2 possibly by co-metabolism. Potentiometric titrations of the compound gave pKa values of 2, 2.6, 5.6 and 10.6. . 1975. Rapid inactivation of glyphosate in the soil. Weed Science 23: 224-28. In greenhouse studies, soil applications of 14C-methyl-labelled glyphosate [N(phosphonomethyl)glycine] were not readily absorbed by corn (Zea mays L. 'Michigan 400') and soybean [Glycine max (L.) Merr. 'Hark']. However, glyphosate available to plants in sand culture was absorbed. Wheat (Triticum aestivum L. Avon) a sensitive bioassay plant, was used to detect the herbicide. Clay loam and muck soil rapidly inactivated 56 kg/ha of glyphosate. Autoclaving of the soil did not prevent the inactivation of glyphosate. In a sandy clay loam soil, application of 56 kg/ha of glyphosate decreased plant growth with increasing pH. Additions of 98 or 196 kg/ha of phosphate to the soil surface decreased glyphosate inactivation in the soil. It is postulated that initial inactivation of glyphosate in soil is by reversible adsorption to clay and organic matter through the phosphonic acid moiety. Stenstrom, J. 1990. Kinetics and effect of moisture content and preincubation of the decomposition of carbon-14-labelled herbicides in soil. Toxicity Assessment 5: 15-28. The kinetics and the influence of soil moisture content (MC) on the mineralization of 14Clabelled linuron were investigated at 15 different MCs in the range from air-dried soil to 100% of the water-holding capacity (WHC). For all MCs used, the data on the liberated amount of 14C (x) were initially described by Eq. (1) x = k1t + a.sbd.after which the data were described by Eq. (2) x = k2t1/2 + b.sbd.where t is time, k1 and k2 are rate constants, and a b are constants. The rate constants k1 and k2 for MCs <100% of WHC were mathematically described by the equation k = 1 + m .cntdot. MC1/2, where k is either of the rate constants, 1 is a constant that can account for a threshold value of MC below which no decomposition occurs, and m is a constant. The validity of this equation for first-order rate constants was tested by using data from the literature. The kinetics of mineralization of 14Clabelled glyphosate were investigated by adding glyphosate after different times of preincubation (tp) of a previously frozen soil. An initial phase of 11 days linear with t [Eq. (1)] was obtained when the herbicide was added immediately after the thawing, after which a phase linear with t1/2 [Eq. (2)] ensued. The length of the initial phase decreased and k1 increased with increasing tp, and for tp .gtoreq. 8 days the initial phase could not be confirmed. The following phase was also unaffected by the preincubation. Thus, the phases represented by Eqs. (1) and (2) are separate, since they can be affected, independently of each other. It is suggested that the initial phase is an induction phase, or lag phase, which reflects the disturbances introduced by handling the soil, and that the second phase is the steady state. (Sweden). Stratton, G. W. 1990. Effects of the herbicide glyphosate on nitrification in four soils from Atlantic Canada. Water, Air and Soil Pollution 51: 373-83. The herbicide glyphosate, supplied as Roundup (Monsanto Canada Inc.), was tested for effects on nitrification in four soils from Atlantic Canada. These included a sandy loam (pH 6.8), two silt loam (pH 6.4 and 5.8) agricultural soils and a clay loam forest soil (pH 3.5). Glyphosate was tested at normal field exposure rates (FR) and levels up to 200 times higher. FR values ranged from 19.83 to 29.26 ppm (.m.g glyphosate/g soil). Glyphosate had no deleterious effects on nitrification in any soil 153

108.

109.

110.

111.

Plant and Soil Residues when tested at FR concentrations. In the sandy loam nitrification was significantly stimulated at a glyphosate level 50 times higher than FR. With this soil and one of the silt loam soils (pH 6.4) glyphosate levels of 100 times FR and higher were required for a significant inhibition of nitrification. With the other silt loam soil (pH 5.8) glyphosate significantly inhibited nitrification at concentrations 10 times FR and higher. Nitrification in the acidic forest soil was very low and accurate toxicity data could not be obtained. The EC50 of glyphosate towards nitrification in soil ranged from 1435 to 2920 ppm, which correspond to exposure levels from 67 to 150 times higher than recommended field application rates. The use of glyphosate in agriculture and forestry should have no toxic effects on nitrification in soil. 112. Stratton, G. W., and K. E. Stewart. 1991. Effects of the herbicide glyphosate on nitrogen cycling in an acid forest soil. Water, Air, and Soil Pollution 60 : 231-47 . The herbicide glyphosate was sprayed aerially on a section of conifer forest in Atlantic Canada that had been previously clearcut and reforested. Glyphosate was then tested for effects on ammonification, nitrification, and denitrification for a period of 8 months by comparing microbial activity in treated and untreated zones of the clay loam forest soil and the overlying decomposing litter, both with a pH of 3.8. With ammonification, there was generally a stimulation of activity in both the forest litter (FL) and forest soil (FS) that had been exposed to glyphosate during spraying. Nitrification rates in FL and FS were very low and glyphosate had no appreciable stimulatory or inhibitory effect on nitrification. Although glyphosate stimulated denitrification in a few instances, it generally had no significant effect on denitrification activity in FL and FS exposed during spraying. With all processes microbial activity in FL was significantly greater than that in FS. Laboratory bioassays were also performed with FL and FS, as well as two silt loam (pH 5.8 and 6.4) and one sandy loam (pH 6.8) agricultural soils, using glyphosate concentrations up to 200 times higher than field application rates. With ammonification and denitrification, glyphosate generally stimulated activity at all levels tested and in all soil used. Glyphosate stimulated ammonification by 50% at concentrations ranging from 140 to 550 mg/g for the soils and >4000 mg/g for FL. With denitrification, the corresponding herbicide levels were approximately 2250 mg/g for FS, >10,000 for FL, and 450 for an agricultural soil. With nitrification, it was estimated that glyphosate concentrations greater than 1000 to 2000 mg/g would be required to cause a 50% inhibition of activity. The careful use of glyphosate in forestry should have no toxic effects on N cycling in soils. Thomas, M. W., B. M. Judy, W. R. Lower, G. F. Krause, and W. W. Sutton. 1989. Time-dependent toxicity assessment of herbicide contaminated soil using the green alga Selenastrum capricornutum. In: Plants for Toxicity Assessment. W. Gorsuch, J. W. Lower, W. R. Wang, (editors), pp. 235-54. See Aquatic Invertebrates and Algae Section. Thompson, D. G., J. E. Cowell, R. J. Daniels, B. Staznik, and L. M. MacDonald. 1989. Liquid chromatographic method for quantitation of glyphosate and metabolite residues in organic and mineral soils, stream sediments, and hardwood foliage. Journal of the Association of Official Analytical Chemists 72: 355-60. A liquid chromatographic method for determining glyphosate (GLYPH) and its major metabolite aminomethylphosphoinc acid (AMPA) in various environmental substrates is described. Ion-exchange column chromatography is coupled with post-column ninhydrin derivatization and absorbance detection at 570 nm. Use of a valve-switching technique allowed quantitation of both analyses in a single chromatographic run and eliminated slow-eluting, coextrated interferences. The method was successfully used to quantitate GLYPH and AMPA in organic and mineral soils, stream sediments and foliage of 2 hardwood brush species. Mean recovery efficiencies for GLYPH as determined from fortified blank field samples were as follows: bottom sediment 84%, suspended sediment 66%, organic soils 79%, mineral soils 73%, alder leaf litter 81%, salmonberry leaf litter 84%, and artificial deposit collectors 87%. Precision for GLYPH determination was good with less than 14% coefficient of variation on mean recovery for all substrates. Limits of detection were lowest for sediments (0.01 g/g dry mass) and highest for foliage substrates (0.10 g/g dry mass). Using this system, 6 samples/person/day were routinely analyzed.

113.

114.

154

Plant and Soil Residues 115. Thompson, D. G., D. G. Pitt, P. T. Buscarini, B. Staznik, D. R. Thomas, and E. G. Kettela. 1994. Initial deposits and persistence of forest herbicide residues in sugar maple (Acer saccharum) foliage. Canadian Journal of Forest Research 24: 2251-62. Initial deposition and subsequent fate of herbicide residues in sugar maple (Acer saccharum Marsh.) foliage were quantified following application of three different formulations of glyphosate (VISION, TOUCHDOWN, MON14420) and one formulation of triclopyr ester (RELEASE) in a comparative field study. Maximum initial residues were 529, 773, 777, and 1630 mg of acid equivalent per kilogram dry mass, respectively. Initial foliar residues were dependent upon application rate (r2=0.63 to 0.87) and increased by a similar factor (233 to 313 mg kg-1) for each kilogram per hectare applied, irrespective of formulation type. Foliar residues dissipated following a negative exponential pattern with time, rates of which varied with initial concentration. Mean times to 50% dissipation were 2 days for all glyphosate formulations, 1.5 days for triclopyr ester, and 4 days for triclopyr acid. Mean times to 90% dissipation were <16 days for glyphosate formulations, 9 days for triclopyr ester, and 33 days for triclopyr acid. Multivariate analyses of intercept and rate parameter estimates indicated significant (p = 0.02) differences in dissipation patterns among treatments. Orthogonal contrasts confirmed a priori hypotheses that glyphosate residue dissipation was independent of the salt formulation applied, and that triclopyr ester dissipated faster than either glyphosate (p = 0.004) or triclopyr acid residues (p = 0.07). Results are considered in terms of the exposure and resultant potential toxicity to forest songbirds inhabiting or foraging in treated hardwood canopies. Thompson, D. G., D. G. Pitt, B. Staznik, N. J. Payne, D. Jaipersaid, R. A. Lautenschlager, and F. W. Bell. 1997. On-target deposit and vertical distribution of aerially released herbicides. Forestry Chronicle 73: (In press) . As a component of the Fallingsnow Ecosystem Project, glyphosate and triclopyr herbicides (Vision, Release) were each applied to four experimental spray plots at nominal rates of 1.5 and 1.9 kg a.e. ha-1 respectively. Empirical studies were undertaken on these plots with the objectives of: (a) quantifying mean on-target deposit and variability (b) assessing the vertical distribution of active ingredient deposits through the vegetative complex and (c) comparing herbicide deposit estimates on excised natural foliage with those on proximal 2-dimensional (2D) and 3-dimensional (3D) collectors. Experimental conditions were representative of difficult aerial application scenarios since the spray plots were small (4.9 to 10.4 ha), with irregular boundaries of mature timber, and in some cases substantial topographical relief. Deposit analysis confirmed that in some circumstances locations well within target areas were missed completely owing to inappropriate track spacing or swath offset. Excluding these points from the data analysis, results demonstrated overall mean deposition (mean SE) of glyphosate and triclopyr on aspen foliage equating to 68.45 6.13 and 50.28 6.01% of the nominal application rates (1.5 and 1.9 kg/ha), respectively. A high degree of variation in deposit both within and between plots demonstrate that variation in operational parameters (e.g. track spacing, offset, release height and aircraft speed) as influenced by local site factors (e.g. proximity of standing timber, topographical relief) can be important determinants in uniformity and accuracy of herbicide deposit. A consistent trend (P < 0.001) in the deposition profile through tiered vegetative canopies was observed, with greatest impingement of the spray in the upper target canopy as noted above, and average 25% and 12% in the shrub and ground-level tiers respectively. Results suggest that for sites characterized by complex canopies, differential vertical deposition may be an important factor constraining the potential use of lower herbicide application rates, particularly where shrub or groundcover species are important competitors. In contrast, given that only a small portion of the spray cloud penetrates and impinges in the lower vegetative tiers, animals foraging or living therein may receive substantially reduced exposures, mitigating against any potential direct effects. In general, poor correlations (r = 0.22 to 0.78) in deposit estimates based on either two-dimensional or threedimensional artificial collectors as compared to excised natural foliage were observed. Significant differences (P < 0.05) also were detected among deposit estimates with no consistent trend in relation to herbicide treatment, sampler type or sampling height. These comparisons suggest that none of the artificial collector types tested accurately or consistently estimated true foliar deposit. Thompson, D.G., D. G. Pitt, T. Buscarini, B. Staznik, D. R. Thomas, and E. G. Kettela. 1995. Initial deposits and persistence of forest herbicide residues in sugar maple (Acer saccharum)

116.

117.

155

Plant and Soil Residues foliage. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand, 20-24 March. R.E. Gaskin, J.A. Zabkiewicz (editors), FRI Bulletin No. 192. pp. 313-315. Initial deposition and subsequent fate of herbicide residues in sugar maple (Acer saccharum Marsh.) foliage were studied following applications of three different formulations of glyphosate and one formulation of triclopyr ester in a comparative field study. Results indicated that initial foliar residues were highly correlated to application rates, increased by a similar factor per unit application rate irrespective of formulation type, and varied substantially in relation to hardwood crown volume index. Dissipation of glyphosate foliar residues followed an exponential decline function, with time to 50% dissipation (DT50) averaging 2 days irrespective of the formulation tested. The general equivalence in foliar impingement and persistence among the isopropylamine, monoammonium and trimethylsulfonium salt formulations of glyphosate tested suggest a similar pattern of behaviour in or on target plant species. The relatively rapid dissipation of triclopyr ester (average DT50=1.5 days), as compared to triclopyr acid dissipation rates (average DT50=4.0 days), suggests poor uptake and translocation of the latter and provides a plausible explanation for the comparatively poor efficacy observed for triclopyr under the conditions of this study. 118. Torstensson, L. 1985. Behaviour of glyphosate in soils and its degradation. In: The Herbicide Glyphosate. Edited by E. Grossbard, and D. Atkinson. London: Butterworths. pp. 137-49. Glyphosate is rapidly adsorbed on soil. Adsorption occurs through the phosphoric acid moiety that competes for binding sites with inorganic phosphates. The extent of adsorption is therefore correlated with the unoccupied phosphate sorption capacity of the soil. Content of organic matter and soil pH have little effect on adsorption, but addition of phosphate decreases adsorption on soils with unoccupied phosphate sorption capacity. Adsorption of glyphosate is reversible. Glyphosate is practically immobile in soils. The mobility is slightly increased at high pH and at high levels of inorganic phosphate in the soil. Glyphosate may, to a minor extent, be degraded by chemical reactions, but the main route of degradation is by microbial activities. Degradation occurs without any lag phase and seems to be a co-metabolic process. It occurs under both aerobic and anaerobic conditions. The principal degradation product of glyphosate is aminomethylphosphonic acid (AMPA). AMPA is also biologically degradable. The rate of degradation of AMPA is often slower than for glyphosate, probably because of stronger adsorption. At normal application rates of glyphosate, Nnitrosoglyphosate is not found in the soil. Rates of glyphosate degradation vary considerably between different soils. Half-lives, ranging from a few days to several months or years, have been detected. The rate of degradation cannot be correlated with a single soil factor, but with the general microbial activity of soils. This activity is an expression of the influence of all soil factors on the microorganisms and is often fairly well correlated with mineralization of the herbicide. The rapid disappearance of herbicidal activity when glyphosate is applied to soil is not due to a single factor. Inactivation through adsorption is of importance. Volatilization does not occur and leaching is practically negligible. Disappearance through degradation is often slow. When applied to plant roots, glyphosate has a low intrinsic toxicity, which is of importance. All factors together create the behaviour of glyphosate in soil. Our present knowledge of glyphosate behaviour and degradation gives no reason to suppose that the herbicide may cause any unexpected damage after application to the soil or elsewhere in the environment. (Sweden). Torstensson, N. T. L., and A. Aamisepp. 1977. Detoxification of glyphosate in soil. Weed Research 17: 209-12. Detoxification of glyphosate [N(phosphonomethyl)glycine) in nonsterile and autoclaved soils was followed by bioassay with wheat. Comparisons were made with detoxification of MCPA [2-methyl4-chlorophenoxyacetic acid] under similar conditions followed by bioassay with spring rape. The well known pattern for microbial metabolism of MCPA with a lag phase preceding the rapid degradation was shown. The initial rapid inactivation of glyphosate is by adsorption, but the results also clearly indicate that the further disappearance of activity depends mainly on microbial degradation. Glyphosate does not seem to sustain microbial growth, which indicates that it is degraded by cometabolism. In autoclaved soil the possibility of a slight chemical degradation or an adsorption that becomes stronger with time could not be excluded. (Sweden).

119.

156

Plant and Soil Residues 120. Torstensson, N. T. L., L. N. Lundgren, and J. Stenstrom. 1989. Influence of climatic and edaphic factors on persistence of glyphosate and 2,4-D in forest soils. Ectotoxicology and Environmental Safety 18: 230-239. Persistence in soil of the two herbicides glyphosate and 2,4-D was investigated after application for brush control in conifer reforestation areas. Field experiments were carried out at five sites in southern Sweden and six in northern Sweden. Initially, glyphosate disappeared faster in northern soils than in southern soils. This was probably a result of the higher biological activity in the northern soils. However, small amounts of glyphosate were detected in the northern area long after all traces had disappeared in the south, presumably because of the long period during which the soil remained frozen in the northern area and because of the slow release of vegetation-bound herbicide. One metabolite of glyphosate, aminomethylphosphonic acid (AMPA), persisted longer than glyphosate itself. After 2 years, 8% of the theoretical amount was found in the northern area and, after 1 year, 1% in the southern area. 2,4-D disappeared rapidly from all sites, although minor amounts persisted for several years, probably because of slow release from vegetation-bound residues. (Sweden). Tropea, M., G. Fisichella, and A. Longo. 1979. Influence of glyphosate on CO2 evolution in some typical soils of eastern Sicily. Tecnica Agricola 31, no. 13: 11 . (Italy). Undabeytia, T., M. V. Cheshire, and D. McPhail. 1996. Interaction of the herbicide glyphosate with copper in humic complexes. Chemosphere 32, no. 7: 1245-50. Glyphosate is a herbicide which is considered to be inacvtivated in soil by bonding to clays and organic matter, possibly through metal bridges. In this paper it has been shown that copper in innersphere complexes with humic substances is complexed by the herbicide. After hydrolysis of the humic acid fraction with water followed by 6M HCl, the glyphosate was unable to react with copper in the humic-copper complex to form a glyphosate complex probably because the copper is more strongly bound on sites previously occupied by iron. (Spain). Uotila, M., G. Gullner, and T. Komives. 1995. Induction of glutathione S-transferase activity and glutathione level in plants exposed to glyphosate. Physiologia Plantarum 93, no. 4: 689-94. Treatment with the herbicide glyphosate led to significantly increased activities of the enzyme glutathione S-transferase (GST, EC 2.5.1.18) in wheat (Triticum aestivum L. cv. Kadett and cv. Satu), pea (Pisum sativum L. cv. Debreceni vilagoszold) and in maize (Zea mays, L. Pioneer 3839 hybrid) tissues. GST activities in wheat seedlings (cv. Kadett) exposed to 960 mu-M glyphosate for 4 days were ca 6-fold and 3-fold higher in shoots and roots, respectively, than in the controls. Glyphosate increased the GST activity to a lesser extent in pea and maize than in wheat. In wheat seedlings (cv. Satu) exposed to 120 mu-M glyphosate gradual increases in the content of non-protein thiols were observed. After 7 days exposure to glyphosate the thiol levels rose to about 360% and 220% of the controls in wheat shoots and roots, respectively. The elevation of thiol content in glyphosate-treated plants was shown to be primarily due to increases of glutathione level. These results suggest that the enhanced glutathione metabolism may have a role in the mode of action or degradation of this herbicide. Wan, M. T. K. 1986. The persistence of glyphosate and its metabolite amino-methyl-phosphonic acid in some coastal British Columbia streams. Environment Canada, Environmental Protection Service, Regional program, Pacific and Yukon Region. Report 85-01. See Water Quality Section. Wardle, D. A., and D. Parkinson. 1990. Effects of three herbicides on soil microbial biomass and activity. Plant Soil 122: 21-28. Three post-emergence herbicides (2,4-D, picloram and glyphosate) were applied to samples of an Alberta (Canada) agricultural soil at concentrations of 0, 2, 20, and 200 m.g/g. The effects of these chemicals on certain microbial variables was monitored over 27 days. All herbicides caused enhancement of basal respiration but only for 9 days following application, and only for concentrations of 200 m.g/g. Substrate-induced respiration was temporarily depressed by 200 m.g/g picloram and 2,4-D and briefly enhanced by 200 m.g/g glyphosate. It is concluded that because changes in 157

121.

122.

123.

124.

125.

Plant and Soil Residues microbial variables only occurred at herbicide concentrations of much higher than that which occurs following field application, the side-effects of these chemicals is probably of little ecological significance. 126. . 1990. Influence of the herbicide glyphosate on soil microbial community structure. Plant Soil 122: 29-38. The side effects of glyphosate on the soil microflora were monitored by applying a range of glyphosate concentrations (0, 2, 20, and 200 m.g/g herbicide) to incubated soil samples, and following changes in various microbial groups over 27 days. Bacterial propagule numbers were temporarily enhanced by 20 m.g/g and 200 m.g/g glyphosate, while actinomycete and fungal propagule numbers were unaffected by glyphosate. The frequency of three fungal species on organic particles in soil was temporarily enhanced by 200 m.g/g glyphosate, while one was inhibited. One species was temporally enhanced on mineral particles. However, many of these fungi were inhibited by 200 m.g/g glyphosate in pure culture. There was little agreement between species responses to glyphosate in incubated soil samples and in pure culture. . 1991. Relative importance of the effect of 2,4-D, glyphosate and environmental variables on the soil microbial biomass. Plant Soil 134: 209-20. Two post-emergence herbicides (glyphosate and 2,4-D) were applied at field application levels to tilled field plots in a mixed cropping area in south-central Alberta. The effects of these chemicals on certain variables associated with microbial biomass and activity were monitored in these plots (as well as corresponding control plots) for 45 days. Glyphosate did not influence any of the microbial variables tested. Addition of 2,4-D significantly influenced all microbial variables investigated but these effects were transient, being detectable only within the first 1-5 days of herbicide addition. The effects of 2,4-D addition on the microbial variables tested, even when significant, were typically small and probably of little ecological consequence especially when spatial and temporal variation in these variables is taken into account. . 1992. The influence of the herbicide glyphosate on interspecific interactions between four soil fungal species. Mycological Research 96: 180-186. The influence of the herbicide glyphosate on two-way interactions between four fungal species (Mucor hiemalis, Mortierella alpina, Trichoderma harzianum, Arthrinium sphaerospermum) was investigated on agar using direct opposition, volatile inhibitor and non-volatile inhibitor tests, and on barley straw using tests for competitive saprotrophic ability. The results indicated that glyphosate was often capable of shifting the direction of two-species interactions, but there was little agreement between the different tests. Reanalysis of particle colonization data presented earlier demonstrated that glyphosate frequently eliminated negative associations between the same four species, possibly by acting as a resource. It is concluded that interactions between pairs of fungi can be influenced directly by glyphosate (especially at high concentrations) and that this may influence soil fungal community structure. . 1992. Influence of the herbicides 2,4-D and glyphosate on soil microbial biomass and activity: a field experiment. Soil Biology and Biochemistry 24: 185-86. Field trials were conducted at a site in S.-central Alberta to evaluate the effects on soil microbial basal respiration (BR) and substrate-induced respiration (SIR) of the following treatments: manual weeding every 3-6 d vs. no weeding, sowing with barley cv. Klondyke at 180 kg seed/ha vs. no crop, and treating with 4 kg/ha 2,4-D or 5 kg glyphosate vs. no herbicide. BR was unaffected by 2,4-D, except after 4 d in barley-sown non-weeded plots, and after 291 d in weeded plots with no crop. The effect of glyphosate on BR was not significant in any plot. 2,4-D did not exert any consistent effects on SIR. However, SIR values were significantly reduced (P = 0.01) by glyphosate after 8 d on plots with weeds but without barley plants. It was concluded that both herbicides were capable of reducing microbial activity or biomass within the 1st few d following application, but only in plots with weeds present.

127.

128.

129.

158

Plant and Soil Residues 130. Wigfield, Y. Y., F. Deneault, and J. Fillion. 1994. Residues of glyphosate and its principle metabolite in certain cereals, oilseeds, and pulses grown in Canada, 1990-1992. Bulletin of Environmental Contamination and Toxicology 53, no. 4: 543-47. Because glyphosate is effective as herbicide and provides harvest management benefits, in 1990 questions were raised from Agriculture Canada field inspection staff regarding the potential misuse of the herbicide which at that time was not registered for pre-harvest use on crops. Thus a post-harvest survey was conducted to monitor glyphosate residues in these cereals, oilseeds, and pulses grown during 1990-1992 period to check if the registration uses of glyphosate were being followed. This paper presents the 3-year monitoring results comprising 459 samples of 8 different crops grown in 7 different provinces in Canada. Wuncheng, W. 1994. Rice seed toxicity tests for organic and inorganic substances. Environmental Monitoring and Assessment 29 , no. 2: 101-7 . Plant seed toxicity tests can be used to evaluate hazardous wastes and to assess toxicity of effluents and industrial chemicals. Cultivation of rice [Oryza sativa], as the test species, in plastic seed trays was investigated as a technique for evaluation of toxicity of metal ions and organic herbicides. Variation of test species, methods, exposure duration and other factors may affect results. Results showed that the order of decreasing toxicity of metal of ions for rice was Cu > Ag > Ni > Cd > Cr(VI) > Pb > Zn > Mn > NaF. Test results were similar to those previously reported for lettuce Ag > Ni > Cd, Cu > Cr(VI) > Zn > Mn, millet Cu, Ni > Cd > Cr(VI) > Zn > Mn and ryegrass Cu > Ni > Mn >> Pb > Cd > Zn > Al > Hg > Cr > Fe. The order of decreasing toxicity of organic herbicides was paraquat, 2,4-D >> glyphosate > bromacil.

131.

159

Terrestrial Invertebrates

Terrestrial Invertebrates
1. Asteraki, E. J., C. B. Hanks, and R. O. Clements. 1992. The impact of the chemical removal of the hedge-base flora on the community structure of carabid beetles (Col., Carabidae) and spiders (Araneae) of the field and hedge bottom. Journal of Applied Entomology 113: 398-406. In commercial farming practice herbicides may be used to eradicate or manipulate the flora in the base of hedgerows. The effect of this on the carabid and spider communities was assessed. The carabid and spider communities of a hawthorn hedge alongside a semipermanent sward and of the adjacent part of the field were sampled using pitfall traps. The hedge was divided into three replicates of two herbicide treatments ('Broadshot' and 'Roundup') and a control. The carabid data were ordinated (DECORANA) and showed that both herbicide treatments affected the communities: the spider data showed similar but less marked responses. Both herbicides had important effects on the carabid communities, but, the spider communities were most affected when only the broadleaved forbs were removed. (United Kingdom). Barker, G. M. 1990. Pasture renovation: interaction of vegetation control with slug and insect infestations. Journal of Agricultural Science 115: 195-202. In a field trial at Rukuhia, New Zealand, Lolium perenne and Trifolium repens seed was direct drilled into pasture (a) without herbicide suppression of the resident sward, (b) with banded application of glyphosate or paraquat herbicide at drilling to remove 50% of the resident sward or (c) after complete removal of the grass and weed components of the old sward with glyphosate or paraquat before drilling. These treatments were compared with the untreated old swards. Where the old sward was removed by herbicide before drilling, pests moved on to the drilled seeding rows, but where herbicides were sprayed in bands over the drill rows, the pests remained in or moved into the residual bands of the old sward. Significant beneficial interactions between herbicide use and in-furrow applications of molluscicides and insecticides resulted in reductions in the number of pests on the seedling rows. The influence of vegetation control on the pest burden is discussed in the context of current pasture renovation practices. (New Zealand). Bergvinson, D. J., and J. H. Borden. 1991. Glyphosate-induced changes in the attack success and development of the mountain pine beetle and impact of its natural enemies. Entomologia Experimentalis Et Applicata 60: 203-12. In field studies conducted in British Columbia, Canada, Roundup (glyphosate) administered into the sapwood around the root collar of lodgepole pine trees, Pinus contorta var. latifolia, was investigated as a tool for creating trap trees for the scolytid Dendroctonus ponderosae. D. ponderosae attacked semiochemical-baited, glyphosate-treated trees before attacking baited control trees. Bark samples from treated trees disclosed an increase in survival of D. ponderosae eggs and early instar larvae, enhanced larval development and increased attacks by insect parasitoids and predators. The braconid Coeloides dendroctoni (C. rufovariegatus) parasitized D. ponderosae larvae found at high densities in treated trees 2 months following attack. The dolichopodid predator Medetera aldrichii was found only in treated trees. Increased competition on P.c. var. latifolia by the scolytid Ips pini had a negative impact on D. ponderosae survivorship. The enhanced impact of insect agents apparently offsets any glyphosate-induced gains in survival of D. ponderosae or its development, but did not cause a significant reduction in emergence. Boller, E. F., E. Janser, and C. Potter. 1984. Testing of the sideeffects of herbicides used in viticulture on the common spider mite Tetranychus urticae and the predacious mite Typhlodromus pyri under laboratory and semifield conditions. Zeitschrift Fur Pflanzenkrankheiten Und Pflanzenschutz 91: 561-68. The sideeffects of herbicides used in viticulture on Tetranychus urticae and the predacious Typhlodromus pyri were studied under laboratory and fieldpot experiments in Switzerland in 1983. Mortality rates obtained in the laboratory did not reflect the situation in the field. A relatively simple laboratory test was used to measure mortality, repellency and survival rates. Generally, the herbicides that showed the most intensive repellency in the laboratory (Reglone (diquat), Gramoxone (paraquat), 160

2.

3.

4.

Terrestrial Invertebrates Semparol (a mixture of atrazine, MCPP (mecoprop) and 2,4,5-T) and Caragard combi (a mixture of terbuthylazine and terbumeton)) were most likely to cause population increases of Tetranychus urticae in the field. The herbicides most toxic to both mite species in the laboratory were Roundup (glyphosate) and Basta (glufosinate). (Switzerland). 5. Brust, G. E. 1990. Direct and indirect effects of four herbicides on the activity of Carabid beetles Coleoptera carabidae. Pesticide Science 30: 309-20. Four herbicides, atrazine, simazine, paraquat, and glyphosate were tested for their acute and chronic toxicity as well as repellent effects on five common carabid beetles (Amara sp., Agonum sp., Pterostichus sp., Anisodactylus sp., and Harpalus sp.) in laboratory and greenhouse experiments. These carabid species are potential biocontrol agents and interference with their biology through herbicide applications, either directly or indirectly, could lead to soil pest outbreaks. Short-term field studies also were conducted to verify laboratory and greenhouse results. The four herbicides did not have significant acute or chronic effects on male or female carabid longevity or food consumption during one year after exposure to initial field-rate applications. Only simazine and atrazine had a repellent effect on carabids, which lasted approximately three days in greenhouse studies. Behavioral studies indicate that once carabids, especially smaller ones (<10 mm in length), establish burrows and foraging territories, they tend to remain in these areas and migrate out of them only slowly. There was no toxic or repellent effect of any herbicide in the field. Carabids, instead, apparently responded to the destruction of plant material which provided a less favorable habitat for the larger (>10 mm length) carabids. Glyphosate and paraquat had the greatest effect on carabids by the second week in field studies, with significantly fewer large carabids found in these two treatments than in the control. Large carabids did not return to paraquat- and glyphosate-treated field areas until approximately 28 days after application; consequently, lower rates of predation of early-season lepidopteran pests by these carabid species may occur in no-tillage corn fields that utilize herbicides for weed control. Christian, F. A., R. N. Jackson, and T. M. Tate. 1993. Effect of sublethal concentrations of glyphosate and dalapon on protein and aminotransferase activity in Pseudosuccinea columella. Bulletin of Environmental Contamination and Toxicology 51, no. 5: 703-9. Clements, R. O., B. R. Bently, and E. J. Asteraki. 1990. Origin of frit-fly (Oscinella spp.) attack on newly sown grass, effect of pesticides on frit-fly larvae and benefit of using glyphosate herbicide pre-ploughing. Crop Protection 9: 105-10. Frit-fly larvae (Oscinella spp.) frequently damage newly sown grass. The infestation arises either by larvae from a previous ploughed-in sward migrating up from depth, or by larvae hatching from eggs laid on or near seedlings by adults that fly in. In an experiment at 5 widely separated sites in England, the importance of these 2 sources of larvae for autumn-sown grass was investigated. Plots were reseeded, after ploughing a previous sward, into areas that were untreated or had been treated with glyphosate and left for larvae to perish before ploughing or not. A range of pesticides was used on the reseeded plots to assess the impact of larvae arising by direct oviposition (fly-in) and to evaluate the chemicals. Chlorpyrifos and an experimental synthetic pyrethroid were the most effective of the chemicals used. Fly-in was the only source of larval attack, as none of the old swards that were reseeded contained a significant population of larvae to migrate. A major benefit of using glyphosate pre-ploughing on seedling emergence and herbage yield was revealed. (United Kingdom). Dalby, P. R., G. H. Baker, and S. E. Smith. 1995. Glyphosate, 2,4-DB and dimethoate: Effects on earthworm survival and growth. Soil Biology and Biochemistry 27, no. 12: 1661-62. Our aim was to test the effect of glyphosate (broad spectrum herbicide), 2,4-DB (broadleaf herbicide) and dimethoate (insecticide and acaricide) on the growth and survival of four earthworm species (Aporrectodea trazezoides Duges, A. rosea Savigny, A. caliginosa Savigny and A. longa Ude) in a pasture soil from the Mount Lofty Ranges, South Australia. The site is being used for research into the ecology of the four earthworm species. The biocides are used to control red-legged earth mites and weed control in field experiments at the site, but may also affect earthworms and unintentionally alter the outcome of the experiments. Therefore the effect of these biocides on earthworm growth and survival needs to be tested. (Australia).

6.

7.

8.

161

Terrestrial Invertebrates 9. Eijsackers, H. 1985. Effects of glyphosate on the soil fauna. In: The Herbicide Glyphosate . Edited by E. Grossbard and D. Atkinson, pp. 151-58. Butterworths, London. (*) To summarize this review paper, it must be concluded that experiments carried out in the field so far indicate that the sideeffects of glyphosate on the soil fauna are small or absent. Nevertheless, there are some results of laboratory experiments which need confirmation. The 100% mortality in the experiments with Amblyseius fallacis, for instance, should be verified under more realistic conditions because of the less drastic effects found with related Acarina such as Gamasida. For the two isopod species Philoscia muscorum and Oniscus asellus, there was a distinct decrease in longevity compared with the control animals (31 and 38%, respectively) with application of recommended amounts of glyphosate (5.96 ml/m2). However, the first species was not affected by glyphosate if offered by means of litter from sprayed shrubs. Although, in both experiments, changes in litter fragmentation through spraying of glyphosate were observed, it is difficult to evaluate effects of this herbicide on litter breakdown. The 30% increase in birch litter fragmentation occurring over the period of 6 months after spraying is well within the range of normal fluctuations in litter breakdown in Mull and Mor soil profiles with a high and low biological activity, respectively. However, these differences might also add to the increased litter fragmentation resulting from herbicide application, leading to increased differences in litter breakdown between treated and untreated plots. Moreover, an evaluation of the differences is connected intimately with the total change in vegetation cover caused by the herbicide. Longterm field trials could elucidate this problem. At present, glyphosate does not seem to be very toxic for the soil fauna, although adverse effects have been observed in studies with a number of representatives of the soil fauna. The limited number of observations, however, does not permit generalizations of these conclusions. (The Netherlands). Forschler, B. T., J. N. All, and W. A. Gardner. 1990. Steinernema feltiae activity and infectivity in response to herbicide exposure in aqueous and soil environments. Journal of Invertebrate Pathology 55: 375-79. Laboratory bioassays were carried out to determine the effect of herbicide exposure on the nematode Neoaplectana feltiae (N. carpocapsae). Alachlor, sethoxydim, 2,4-D and glyphosate were tested at 3 concn in Petri dishes including the recommended application rate and one log dose higher and one log dose lower than the recommended rate for each of 3 exposure periods of 24, 48 and 120 h. Alachlor was the most deleterious, reducing dauer activity by 44 to 90% in aqueous solution. Only 19 to 30% of the nematodes exposed to the other 3 herbicides were inactive. Bioassays of nematodes extracted from these suspensions and subsequently washed 3 times in water against final-instar larvae of Galleria mellonella yielded LD50s of 45.2, 1.1, 0.9 and 1.1 nematodes/larva for the alachlor, 2,4-D, glyphosate and sethoxydim treatments, resp. Control treatments with water yielded an LD50 of 2.8 nematodes/larva. Nematode infectivity in soil was reduced by exposure to 2,4-D and alachlor. At the recommended application rate, the LD50s for the alachlor, 2,4-D and control treatments were 26.7, 8.8 and 2.7 nematodes/larva, resp. Fujiie, A., T. Yokoyama, M. Fujikata, M. Sawada, and M. Hasegawa. 1993. Pathogenicity of an entomogenous nematode Steinernema kushidai Mamiya (Nematode: Steinemematidae), on Anomala cuprea (Coleoptera: Scarabaeidae). Japanese Journal of Applied Entomology and Zoology 37: 53-60. Laboratory and field experiments were conducted to study the infectivity and propagation of Steinernema kushidai on Anomala cuprea larvae. The nematode-infective juveniles (JIII) showed higher infectivity for the 3rd instar than the 2nd and 1st instar of A. cuprea. When JIII were inoculated on the 3rd instar, as much as ca. 30,000 JIII were produced per larva on average. The JIII production increased with increased insect size. Survival of JIII was also investigated under various conditions. The survival rate of JIII stored in water exceeded 50% at 15 degrees C after 3 months, and decreased to 50% in a month or less at 4 degree, 10 degree 20 degree and 25 degree C, and to below 10% at 4 degree C. In sterilized soil, nematode infectivity remained high for up to 6 months at 15 degree C. When the JIII were applied once to the soil of a peanut field in mid-April at the rate of ca. 10-6/m2, the infectivity remained high enough to kill A. cuprea larvae for 8 months, but it became very low during winter. The nematode was adversely affected by treatments with insecticides (Diazinon and MPP, 5002,000 ppm), but not with fungicide (Thiophanatemethyl, 500-2,000 ppm) and with herbicides (CAT and

10.

11.

162

Terrestrial Invertebrates glyphosate, 5,000-20,000 ppm). Twenty minutes or longer exposure of the nematode suspension to sunlight resulted in decreased survival of JIII. Metarhizium anisopliae had no effect on the nematode. 12. Gomez, M. A., M. T. Perez, and M. A. Sagardoy. 1989. Effect of successive applications of glyphosate on the aerobic bacteria and microarthropods of a sandy soil. Ciencia Del Suelo 7: 55-61. Repeated applications of glyphosate at approx. 40 litres/ha on a non-cultivated sandy soil did not affect the density of aerobic bacteria and microarthropods in the soil. In laboratory studies 2.38% of the aerobic bacteria survived 1% glyphosate. It was concluded that gram-negative bacteria could possibly use glyphosate as a nutrient source. (Argentina). Gomez, M. A., and M. A. Sagardoy. 1985. Influence of glyphosate herbicide on the microflora and mesofauna of a sandy soil in a semiarid region. Revista Latinoamericana De Microbiologia 27: 351-58. The effect of four applications (0, 2, 4 and 8 L/ha) of the herbicide glyphosate (Nphosphonomethyl glycine) on the total numbers of aerobic bacteria, microarthropods, mites and spring tails were studied periodically during 96 days in a sandy soil located in the semiarid region of the Buenos Aires province, Argentina. The total numbers of the microflora did not change significantly (p = 0.05) neither with time nor with the mentioned treatments. The total numbers of the other groups were unaffected significantly by the herbicide treatments. However, significant variations with time (p < 0.01) took place in the microarthropod population. Furthermore, a highly significant (R = 0.95**) relationship was obtained between the total numbers of arthropods and mites. Soil humidity was shown to have no effect on the total number of the organisms studied. From the results, it may be concluded that glyphosate herbicide, applied even in amounts double that usually recommended, does not produce harmful effects on the microflora and the mesofauna studied. (Argentina). Hawkins, J. W., M. W. Lankester, R. A. Lautenschlager, and F. W. Bell. 1997. Effects of alternative conifer release treatments on terrestrial gastropods in northwestern Ontario. Forestry Chronicle 73: (In press). Changes in terrestrial gastropod species richness and density in regenerating spruce plantations following application of four conifer release treatments including two chemical herbicides (Vision [a.i. glyphosate] and Release [a.i. triclopyr]) cutting by mechanical means (Silvana Selective/Ford Versatile), and motor-manual cutting with brush saws are discussed. Mean gastropod density increased to a maximum of 21 m-2 over the summer of 1993 (pre-treatment) when a total of 27,396 gastropods were collected but remained fairly stable at about 10 m-2 throughout the summer of 1994 (post-treatment) when only 20,199 were collected. This difference between years was not related to the conifer release treatments, since gastropod densities on both treated and control aras were lower in 1994. The decline in 1994 was probably due to decreased population size and detectibility becuase of less rainfall. The lack of difference following treatments was attributed to rapid re-establishment of the herbaceous layer which probably continued to provide favourable conditions for snails and slugs. Gastropod density was higher in a nine-year-old regenerating spruce plantation (15.51.3 m-2) than in a 70-year-old mixedwood forest (9.40.6 m-2) and species richness was also slightly greater (20 spp. vs. 18 spp.) in the plantation. These differences were attributed to the more abundant near-ground vegetation and the greater amount of deciduous litter characterizing the regenerating plantation. Heyd, R. L., R. L. Murray, and L. F. Wilson. 1987. Managing Saratoga spittlebug in pine plantations by suppressing sweetfern. Northern Journal of Applied Forestry 4: 16-17. One Jack pine and 4 red pine plantations in Michigan, with trees 1.5-3.0 m tall, had sweetfern (Comptonia peregrina) that neared or exceeded the threshold for supporting large populations of spittlebugs (Aphrophora saratogensis). Water (control) or aqueous solutions of Roundup (glyphosate) at 1, 2 or 3 quart/acre and Krenite (fosamine) at 1, 2 or 3 gal/acre were applied in autumn 1979. Number of sweetfern stems and the percentage of sweetfern ground cover were recorded annually in summer 1979-82. All herbicide treatments reduced sweetfern stems by 92-100% in 1980 and reduction was still 58-90% in 1982. Canopy cover was 3% in 1980 and less than or equal to 10% in 1982. On control plots, canopy cover increased from 27 to 44% during the study.

13.

14.

15.

163

Terrestrial Invertebrates 16. Hislop, R. G., and R. J. Prokopy. 1981. Integrated management of phytophagous mites in Massachusetts (U.S.A.) apple orchards. 2. Influence of pesticides on the predator Amblyseius fallacis (Acarina: Phytoseiidae) under laboratory and field conditions. Protection Ecology 3: 157-72. Forty orchard spray materials were evaluated in laboratory and/or field trials for toxicity to different strains of Amblyseius fallacis, the most important predator of the spider mites Panonychus ulmi and Tetranychus urticae in Massachusetts apple orchards. In laboratory trials, materials highly toxic to the Bishop strain were methomyl, carbaryl, phosalone, diazinon, demeton, dimethoate, fenvalerate, permethrin, formetanate hydrochloride, ammonium sulfamate, paraquat, and glyphosate. Materials of moderate toxicity were phosphamidon, dicofol, karathane, cyhexatin, and daminozide. Materials of low toxicity were malathion, phosmet, azinphosmethyl, methyl parathion, endosulfan, methoxychlor, propargite, fenbutatinoxide, benomyl, dodine, glyodin, maneb, dikar, thiram, dichlone, captan, ferbam, manebzinc, simazine, dalapon, ethephon, NAA and calcium chloride. Benomyl leaf residues interfered with reproductive capability. In field research plot trials, phosalone, permethrin, and benomyl reduced the numerical response of A. fallacis to prey increase, resulting in spider mite buildup. Endosulfan, azinphosmethyl, phosmet, and methyl parathion had little effect on A. fallacis abundance, resulting in predatorprey ratios favourable to biological control of spider mites. In commercial orchards receiving combinations of phosalone, formetanate hydrochloride, ammonium sulfamate, benomyl, and/or glyodin, A. fallacis was unable to respond maximally to increasing spider mite abundance, necessitating repeated acaricide treatments. In commercial orchards receiving none of these materials, but instead treated with azinphosmethyl, phosmet, endosulfan, captan, and/or dodine, spider mites rarely succeeded the economic injury level, owing in large part to the more favourable A. fallacisspider mite ratios. House, G. J. 1989. Soil arthropods from weed and crop roots of an agroecosystem in a wheatsoyabean-corn rotation: impact of tillage and herbicides. Agriculture, Ecosystems and Environment 25: 233-44. Soil arthropods were collected in 1985 and 1986 from the root systems of weeds, maize and wheat within experimental, low-input agroecosystems located in the coastal plain region of North Carolina. Eight treatments forming a gradient from high (conventional tillage) to low (no-tillage) soil disturbance in combination with herbicide (glyphosate, alachlor, atrazine, linuron) and non-herbicide treatments were established in Oct. 1985 in a wheat/soyabean/maize rotation. Soil arthropod density was consistently higher under weedy, no-tillage treatments than conventional tillage in both 1985 and 1986. Weed root biomass and soil arthropod number were moderately correlated although arthropod number differed widely among individual weed species. Below ground phytophagous arthropod number from the root systems of maize, dogfennel (Eupatorium capillifolium), ragweed (Ambrosia artemisiifolia), lambsquarter (Chenopodium album) and pigweed (Amaranthus retroflexus) did not differ significantly. However, predacious arthropod number from dogfennel, ragweed, lambsquarter and pigweed root systems was consistently higher than phytophagous arthropods, suggesting the potential for employing selected weed species as predator refuges. House, G. J., A. D. Worsham, T. J. Sheets, and R. E. Stinner. 1987. Herbicide effects on soil arthropod dynamics and wheat straw decomposition in a North Carolina no-tillage agroecosystem. Biology and Fertility of Soils 4: 109-14. Herbicide combinations of paraquat, glyphosate, alachlor, linuron, fluazifopbutyl, aciflurofen, and bentazon were investigated for their impact on soil arthropod population dynamics and surface wheat straw decomposition (weight loss) within a North Carolina coastal plain agroecosystem. Herbicides were applied twice (preemergence and mid-bloom) at recommended field rates to soybean no-till planted into wheat residue. Separate measurements were made for surface crop residue and soil-dwelling (0-3 cm depth) arthropods. Decomposition of herbicide (glyphosate) and nonherbicidetreated wheat straw residue was compared using mesh bag techniques. Decay rate constants were estimated for glyphosate and nonherbicide-treated wheat straw residue by fitting a two-component model to the data. Comparison of soil microarthropod numbers from herbicide and nonherbicide treatments showed no consistent trend, suggesting that abiotic factors such as soil temperature and moisture were more significant than herbicide effects in regulating soil microarthropod number and activity. Herbicides had no effect on soil macroarthropod number or activity until late in the season 164

17.

18.

Terrestrial Invertebrates when macroarthropods were most abundant under weedy, no-tillage conditions. Moist soil and litter, low soil temperature, floral diversity, and high weed-seed availability probably enhanced macroarthropod numbers in nonherbicide treatments. Decomposition of nonherbicided, surface crop residues was more rapid than herbicide (glyphosate) treated, indicating that herbicide effects occur at the decomposer as well as producer level of agroecosystems. 19. Hoy, J. B., D. L. Dahlsten, and P. J. Shea. 1982. The effects of 2,4-D and an alternative material (glyphosate) used for brushfield rehabilitation on soil arthropod community and litter decomposition rate., Final Report USAUCB, Research Agreement PSW-81-0026. 34 p. Kegel, B. 1989. Laboratory experiments on the side effects of selected herbicides and insecticides on the larvae of three sympatric Poecilus spp. Coleoptera carabidae. Journal of Applied Entomology 108: 144-55. The side effects of the insecticides Ekamet, Dursban and Hostaquick and the herbicides Fusilade, Simazin, Atrazin, Roundup, MCPA and Tribunil on the larvae of 3 Poecilus species were investigated by the means of laboratory rearing. The 3 insecticides killed all L1-larvae of Poecilus versicolor, P. cupreus and P. lepidus within 16 days. The L2- and L3-larvae of P. versicolor were not affected by Ekamet after 28 days. In the case of Hostaquick the number of surviving larvae increased with the age and size of the tested larval instars. All larvae which came in contact with the insecticides showed signs of severe poisoning already after two days. The mortality due to the investigated herbicides was much less than 50% in all cases. Significant reductions of the numbers of hatched adults were observed only for the herbicide MCPA and the species P. versicolor. P. lepidus, which did not reach pupation under the experimental conditions (15oC), was affected by Fusilade and Atrazin. the time needed for development until the hatching of the adults was found to be significantly longer than normal, when L2-larvae of P. versicolor were tested with the insecticide Ekamet and Hostaquick. No differences were recorded in the experiments with herbicides. Because of the higher susceptibility of the larval stages of Carabid beetles, the described method is useful and discussed as a possible completion to the work of the IOBC/WPRS-group "Pesticides and Beneficial Organisms". (Germany). Leslie, G. W. 1977. The toxicity of some agrochemicals to Pheidol sp. (Hymenoptera: Formicidae) a common ant in natal cane fields. Proceedings of the Annual Congress of South African Sugar Technology Association. 51:21-23. (South Africa). Mansour, F. 1987. Effect of pesticides on spiders occurring on apple and citrus in Israel. Phytoparasitica 15: 43-50. (Israel). Martin, N. A. 1982. The effects of herbicides used on asparagus on the growth rate of the earthworm Allolobophora caliginosa. Proceedings of the 35th New Zealand Weed and Pest Control Conference. pp. 328-331. (New Zealand). Mohamed, A. I., G. A. Nair, H. H. Kassem, and M. Nuruzzaman. 1995. Impact of pesticides on the survival and body mass of the earthworms Aporrectodea caliginosa (Annelida: Oligochaeta). Acta Zoologica Fennica 196: 344-47. Aporrectodea caliginosa is a common earthworm in the farmlands of Benghazi, Libya. Experiments were performed to study the effects of 25, 50, 100, 200 mg pesticides/kg dry soil of methomyl and lebaycid (insecticides), paraquat and glyphosate (herbicides) and remiltine (fungicide) on the survival and body mass of the worm. LC sub(50) (mg/kg) was very low for the insecticides methomyl (48-19) and lebaycid (46-40). However, it was moderate to high for the herbicides paraquat (182-108), glyphosate (246-177) and for the fungicide remiltine (246-163). The highest toxic effects on the worms were evident with insecticides followed by herbicides and fungicide. A decrease in the body mass was discernible after an exposure period of 8 days in all pesticide concentrations examined. The ecological implications of the use of these pesticides on beneficial organisms such as earthworms are discussed in detail. (Libya).

20.

21.

22.

23.

24.

165

Terrestrial Invertebrates 25. Moore, D., R. O. Clements, and M. S. Ridout. 1987. Effects of pasture establishment and renovation techniques on the Hymenopterous parasitoids of Oscinella frit L. and other stemboring Diptera in Ryegrass. Journal of Applied Ecology 23: 871-82. Different levels of attack by Hymenopterous parasitoids occurred in stem-boring larval Diptera in swards which had been re-seeded after conventional cultivation or by direct drilling or which had been improved by the addition of nitrogen fertilizer (England, UK). Numbers of the major parasitoid of established grass, Chasmodon apterus Nees (a wingless Braconid), were greatly reduced by ploughing and re-seeding, but were reduced to a lesser extent by chemical desiccation with the herbicide glyphosate and direct drilling. Other (Alate) parasitoids showed greater prevalence in seeding grass than established grass. Four months after sowing, over 40% of stem-boring larvae were parasitized. Changes in host population numbers were correlated with the initial level of parasitoid attack. (United Kingdom). Moore, D., M. S. Ridout, and R. O. Clements. 1988. Mortality of Oscinella spp. due to parasitism in insecticide treated and untreated ryegrass reseeds. Journal of Applied Entomology 105: 154-59. During field studies carried out in southern England in 1983, a large proportion of stem-boring larvae of Oscinella spp. attacking Italian ryegrass (Lolium multiflorum) which had been direct drilled, were parasitized by Hymenoptera. The test plots were sprayed with chloropyrifos at 720 g a.i./ha or dimethoate at 350 g a.i./ha or left untreated. The wingless braconid Chasmodon apterus was the most common parasitoid. Chemical destruction of the old sward using the herbicide glyphosate prior to reseeding did not significantly reduce the proportion of stem-boring larvae parasitized by C. apterus. Numbers of parasitized host larvae decreased to a greater extent than the number of unparasitized larvae between 2 sampling dates (12 October and 10 November). Levels of premature mortality among parasitized larvae were higher in areas where initial parasitism was high compared with those from areas of low initial parasitism. The level of parasitism was significantly higher in the insecticidefree plots on the 1st sampling date but not on the 2nd. (United Kingdom). Naguib, N. A., F. K. El-Baz, and R. El-Masry. 1987. Effect of glyphosate on the amino acids and mineral contents of Vicia faba and Orobanche crenata and its influence on the host-parasite relationship. Egyptian Journal of Physiological Science 12: 91-102. A pot experiment was conducted over two successive years to study the effectiveness of one, two or three glyphosate application(s) on Vicia faba L. for controlling Orobanche crenata forsk infestation. Single spraying treatment did not prevent Orobanche infestation. However, two successive spraying treatments were efficient to prevent infestation. Both Orobanche infestation and glyphosate spraying treatments, caused pronounced decrease in the total amino acid content in the leaves of the host plant, however, the decrease induced by the parasite was higher than that induced by glyphosate. On the other hand, the total amino acid content in the roots of the host was not affected by Orobanche infestation, meanwhile, sharp decrease occurred when the host plant was treated with glyphosate. The data suggest that the decrease in the amino acid phenylalanine in the roots of Vicia faba plants induced by glyphosate may play a role in preventing Orobanche infestation. Iron was accumulated in the roots of the host as a result of Orobanche infestation and glyphosate application. (Egypt). Nair, G. A., A. I. Mohamed, and K. C. Bhuyanb. 1995. Comparative effects of chemical pesticides on survival, body mass and respiration of the pulmonate slugs, Milax rusticus (Millet, 1843) and Milax sowerbyi (Ferussac, 1823) (Mollusca: Milacidae). Journal of African Zoology 109, no. 2: 141-49. Laboratory observations, over a period of 144 hours, were made to study the effects of pesticides methidathion (organophosphate), sevin (carbamate), fenvalerate (pyrethroid), glyphosate (herbicide) and ethirimol (fungicide) on the survival, body mass and respiration of the slugs Milax rusticus (Millet, 1843) and Milax sowerbyi (Ferussac, 1823). These pesticides are in common use in the farms of Benghazi and its neighbourhood. Survival was adversely affected for M. rusticus treated with methidathion and sevin wherease M. sowerbyi was more susceptible to methidathion, glyphosate and ethirimol. An increase in the body mass of the control and the pesticide-treated slugs over time

26.

27.

28.

166

Terrestrial Invertebrates was evident in both species. Significant differences in moisture content existed between control and pesticide-treated slugs and also between the pesticide-treated species. The duration of exposure of these slugs in most of the pesticides showed significant increase in their rate of oxygen consumption. The ecological implications of the findings are discussed. (Libya). 29. Patra, P. C., and S. Ray. 1987. Population growth patterns of Helicotylenchus dihystera and Hoplolaimus indicus in weedicide treated groundnut plots. Indian Journal of Nematology 17: 218-20. Populations of H. dihystera and H. indicus were selectively influenced by the 3 herbicides, oxadiazon 25 EC, fluazifop 25 EC and glyphosate 36 EC (each at 0.75 or 1.0 kg a.i./ha singly and 0.75 kg a.i./ha + 0.25 kg a.i./ha 80% sodium salt of 2,4-D in combination). The first 2 suppressed H. dihystera as well as monocot weed growth but did not significantly affect H. indicus. Both the nematode species were directly influenced by the growth stage of the groundnut (Arachis hypogaea) crop, with peak populations being attained around late flowering. Along with several monocot weeds, H. dihystera was either maintained or increased but never decreased, in glyphosate-treated plots, suggesting that unlike H. indicus it finds alternative hosts in some of the weeds and is not affected as much by the declining A. hypogaea crop. (India). Pawlizki, K. H. 1987. Effects of herbicides used in orchards and vineyards on man and the environment. Gesude Pflanzen 39: 486-96. See Plant and Soil Residues. Quimby, P. C., Jr., and K. E. Frick. 1985. Evaluation of herbicide-coated larvae of Bactra verutana Lepidoptera tortricidae to control nutsedges Cyperus rotundus and Cyperus esculentus. Entomophaga 30: 287-92. Feeding by 10-day old mid-instar larvae of the native moth Bactra verutana zeller, coated with solutions of Bentazon (3-isopropyl-1-H-3-benzothiadiazin-4(3H)-one 2,2-dioxide) or glyphosate (N(phosphonomethyl)glycine), reduced the dry weight of purple (Cyperus rotundus L.) or yellow nutsedge (C. esculentus L.) an average 25% more than did uncoated larvae. The herbicide-coated larvae reduced dry weight by 71% for 4-week yellow nutsedge and by 80% for 2-week purple nutsedge plants within 15 days after release. This integrated approach was apparently not practical, however, because larvae randomly damaged the non-target crops, cotton (Gossypium hirsutum L.) and turnip (Brassica rapa L.). The more mature larvae killed about 3% of the cotton and 14% of the turnip plants. Newlyemerged larvae were highly sensitive to the herbicides. Mature larvae are less host-specific than are newly-emerged larvae. Therefore, the use of repeated broadcasts of newly-emerged B. verutana larvae is still the best method known for using insects to control nutsedge. Robinson, A. F., C. C. Orr, and J. R. Abernathy. 1977. Influence of Nothanquina phyllobia on silverleaf nightshade. Proceedings of the 30th Annual Meeting of the Southern Weed Science Society. p. 142. Roorda, F. A., G. G. M. Schulten, and A. H. Pieterse. 1978. The susceptibility of Orthogalumna terebrantis Wallwork (Acarina: Galumnidae) to various pesticides. Proceedings of the 5th EWRS International Symposium on Aquatic Weeds. pp. 375-81. Orthogalumna terebrantis, a mite which has potential for integrated control of Eichhornia crassipes, was tested for its direct susceptibility to 2,4-D, glyphosate, diquat and paraquat. Only diquat was markedly toxic to adults. In addition, direct toxicity of certain insecticides was studied. These included insecticides which are used for the control of disease-carrying insects in marshes and of agricultural pests in the near vicinity of water bodies where E. crassipes occurs. Susceptibility of adults and immatures to these compounds varied considerably, diazinon and endosulfan giving the highest mortality. (The Netherlands).

30.

31.

32.

33.

167

Terrestrial Invertebrates 34. Ross, D. W., C. W. Berisford, and J. F. Godbee, Jr. 1990. Pine tip moth, Rhyacionia spp., response to herbaceous vegetation control in an intensively site-prepared loblolly pine plantation. Forest Science 36: 1105-18. Herbicide release treatments were applied to an operationally regenerated loblolly pine (Pinus taeda) plantation in Georgia, USA, to assess the influence of vegetation control on damage by R. frustrana. Treatments included: banded herbicide (sulfometuron, hexazinone and glyphosate) applications (1.5-m wide bands centered over rows of seedlings) for 1 or 2 years after planting; broadcast herbicide applications (total area covered) for 1 or 2 years after planting; and an untreated control. All of the herbicide treatments significantly reduced the amount of herbaceous vegetation and improved the survival and growth of planted pines. By the end of the third growing season, pine survival in the control plots (67%) was significantly less than survival in all of the herbicide-treated plots (more than or equal to 92%). Trees in herbicide-treated plots had more tip moth damage than trees in the control plot during the first two growing seasons after planting, but by the third growing season damage to trees in the control plots was equal to that in the herbicide-treated plots. The method or number of herbicide applications did not have a significant effect on tip moth infestations. Pine growth gains resulting from reduced vegetative competition were more than enough to compensate for greater tip moth damage following herbicide treatments. However, increased tip moth damage still reduced potential growth gains. Two herbicide applications did not significantly improve pine survival or growth compared to one application, but the broadcast treatments produced slightly larger trees than the banded treatments by the end of the third growing season. Overall, the results indicate that, when assessing the feasibility of R. frustrana control or the optimum time to apply control treatments, the effect of other silvicultural treatments on tip moth population dynamics should be considered. Saly, A. 1991. Evaluation of herbicide influence on the edaphon in vineyards by means of freeliving nematodes. Polish Ecological Studies 15: 47-54. A method of bioindication in soils treated with herbicides based on soil nematodes is described. The composition and biomass of nematode communities was evaluated in 25 ml soil samples from vineyards in Czechoslovakia treated once or repeatedly during 4-8 years at recommended rates. Repeated soil treatment was more harmful to nematode communities than a single application. The species of nematodes found, their numbers and biomass and the effects of Semparol 1167 (atrazine + mecoprop + 2,4,5-T), Prefix G (chlorthiamid), Herbex (simazine), Caragard Combi (terbumeton + terbuthylazine), Ustinex Special (amitrol + 2,4-D + diuron), Gesatop 50, and Roundup are tabulated. Gesatop 50 (48.5% simazine) and Roundup (glyphosate 360 g/litre) increased nematode numbers and are recommended for preserving the soil properties. (Czechoslovakia). Saly, A., and P. Ragala. 1984. Free living nematodes bioindicators of the effects of chemicals on the soil fauna. SB Uvtiz (Ustav Vedeckotech Inf Zemed) Ochr Rostl 20: 15-21. Soil nematodes are bioindicators of pesticide and fertilizer effects on edaphon. In the vineyard soil treated with Roundup herbicide, 38 nematode species were identified. The application rate of 12 L/ha stimulated the nematode community numbers; 2.5 times higher rate reduced them on the average to 63.8% of the control. Nematodes reacted to the herbicide rates and to the sampling terms by changes in population levels, community composition and weight of their biomass and bound energy value. (Czechoslovakia). Schuytema, G. S., A. V. Nebeker, and W. L. Griffis. 1994. Effects of dietary exposure to forest pesticides on the brown garden snail Helix aspersa Mueller. Archives of Environmental Contamination and Toxicology 26, no. 1: 23-28. Brown garden snails, Helix aspersa, were fed prepared diets with 12 pesticides used in forest spraying practices where endangered arboreal and terrestrial snails may be at risk. Acephate, atrazine, glyphosate, hexazinone, and picloram were not lethal at concentrations of 5,000 mg/kg in 14-day screening tests. The remaining seven pesticides, lethal to 13-100% of the tested snails at 5,000 mg/kg, were evaluated in 10-day definitive feeding sites. Azinphosmethyl (Guthion) and aminocarb were the most, with 10-day LS50s of 188 and 313 mg/kg, respectively. Paraquat, trichlorfon and fenitrothion had 10-day LC50s of 659, 664, and 7,058 mg/kg, respectively. Avoidance of pesticide-containing foods occurred, e.g., 10-day LC50s of gt 10,000 mg/kg for carbaryl and methyl parathion. Significant decreases (p < 0.05) in snail weight (total, shell-only, body-only) or shell diameter were accompanied 168

35.

36.

37.

Terrestrial Invertebrates by a significant decrease in the amount of food consumed/snail/day. Concentrations of pesticide in tissues were measured in snails exposed to atrazine and azinphosmethyl; there was no bioaccumulation. 38. Vainio, A., and H. Hokkanen. 1990. Side-effects of pesticides on the entomophagous nematode Steinernema feltiae, and the entomopathogenic fungi Metarhizium anisopliae and Beauveria bassiana in the laboratory. Proceedings and abstracts, Vth International Colloquim on Invertebrate Pathology and Microbial Control. p. 334. Studies on the toxicity of pesticides to Steinernema feltiae (Neoaplectana carpocapsae), Metarhizium anisopliae and Beauveria bassiana are summarized. Bentazone, ioxynil, hexaconazole, cyromazine and buprofezin did not affect N. feltiae, but quizalofop-ethyl, tralkoxydim, sulfur and potassium soap were toxic. Pirimicarb, cypermethrin, diazinon, simazine and metalaxyl plus mancozeb did not affect either fungus, but glyphosate, dimethoate, MCPA, vinclozolin, trifluralin, thiram and propiconazole inhibited at least one species. (Finland). Vythilingam, I., and T. H. Chua. 1983. The effects of herbicides on soil Collembola in Oil Palm Estates. Tropical Ecology 23: 165-72. The effect of 2 herbicides, Roundup and Paracol on soil Collembola was investigated in an oil palm estate (Malaysia) from July 1977-June 1978. Compared to the control plot, the Collembola population increased slightly (p = 0.05) in the plots treated with Roundup at 11 oz/acre; Roundup at 15 oz/acre and Paracol at 2 pt/acre significantly reduced the Collembolan population, with Paracol having a significantly greater effect (p = 0.05). Other factors like moisture content, presence of cover crops, food supply and predator population may also indirectly affect the Collembolan population. (Malaysia). Wright, M. A., D. A. Kendall, and B. D. Smith. 1985. Toxicity of paraquat, paraquat + diquat and glyphosate to the cereal aphid Rhopalosiphum padi. In Tests of agrochemicals and cultivars, London. Association of Applied Biologists. pp. 8-9. Recent experiments suggest that spread of barley yellow dwarf virus (BYDV) by cereal aphid vectors moving from volunteers, grass weeds and grass leys to following cereal crops can be controlled as a side effect of herbicide use. Experiments were done to find if paraquat, paraquat + diquat and glyphosate are toxic to Rhopalosiphum padi which is one of the most important vectors of BYDV between cereals and grasses. Aphids were rapidly killed by paraquat or paraquat + diquat when sprayed either on or off plants, but not by glyphosate. They were also killed by paraquat and paraquat + diquat when placed on plants after sprays had been allowed to dry. R. padi characteristically feeds near the bases of cereal plants and aphids survived for 17 hours when only the tops of the plants received paraquat or paraquat + diquat. Where only the bases of plants had received spray, some aphids moved upwards away from the treated tissue but most died. Paraquat and paraquat + diquat killed aphids on plants when used in concentrations equivalent to 200 g a.i./ha (i.e.) 1/4 of the recommended field dose for herbicide activity. When fed to aphids in sucrose solutions through Parafilm membranes both paraquat and paraquat + diquat killed aphids at concentration equivalents of 100-400 g a.i./ha. At the highest concentration mortality was less because aphids were deterred from feeding. Glyphosate was not toxic when used in this way. (United Kingdom). Ye, Z., and Y. Lu. 1991. Toxicity and residue toxicity of the pesticides on the larvae of Bombyx mori. Canye Kexue 17: 84-89. (China). Yokoyama, V. Y., and J. Pritchard. 1984. Effect of pesticides on mortality, fecundity, and egg viability of Geocoris pallens (Hemiptera: Lygaeidae). Journal of Economic Entomology 77: 876-79. Sublethal doses of the insecticides carbaryl, methidathion and methomyl, the acarides dicofol and propargite, the herbicide glyphosate, the defoliant DEF (s,s,s-tributyl phosphorotrithioate) and the fungicide sulfur were tested as cotton leaf residues on Geocortis pallens, a major predator in Californian cotton. No pesticide except propargite significantly increased adult mortality 25-192 hours after a 24-hour acute exposure; neither did they have any detrimental effect on female mortality, fecundity or egg viability. This suggests that deleterious longterm effects in predator populations would 169

39.

40.

41.

42.

Terrestrial Invertebrates not occur if insects survived initial treatments. Males were more susceptible to methidathion and methomyl and females were more susceptible to propargite. Females exposed to DEF laid more eggs. Potential beneficial effects of chemicals on predator populations should be considered in the selection of pesticides for pest management programs. 43. Zaitseva, V. K. 1981. Specific behavioral reactions of forest red ants to glyphosate treatment. Ekol. Zashch. Lesa 6: 71-75. Broadleaf tree control exposes the beneficial forest red ant (Formica rufa) to the arboricide glyphosate. Spraying anthills at 2-6 kg a.i./ha did not affect ant behaviour or the expansion of anthills. (Soviet Union).

170

Water Quality

Water Quality
1. Anton, F. A., L. M. Cuadra, P. Gutierrez, E. Laborda, and P. Laborda. 1993. Degradation behavior of the pesticides glyphosate and diflubenzuron in water. Bulletin of Environmental Contamination and Toxicology 51, no. 6: 881-88. (Spain). Batty, J. 1991. Safeguarding water supplies. "NTC Workshop." Report No. 22, pp. 28-33. Efforts by the Severn Trent region of the National Rivers Authority (NRA, a regulatory body set up in the UK as a consequence of the 1989 Water Act) to persuade highway and railway authorities to use glyphosate instead of atrazine and simazine for weed control along roadsides and railway lines, resp., and so reduce residues of the triazine herbicides in the surface and groundwaters within its remit, are discussed. The NRA recognizes that it may face some resistance to its proposal from transport authorities and sub-contractors employed by them for weed control treatments, as glyphosate is more expensive and is effective for a shorter duration than atrazine or simazine, and hence warns that it may resort to bringing prosecutions under sections 107 and 108 of the Water Act. (United Kingdom). Beck, A. E. 1987. Persistence of glyphosate residues in Northern Manitoba ponds. Water Standards and Studies. Report no. 87-2. Manitoba Environment and Workplace Safety and Health Environmental Management Services. The purpose of the present study was to accurately determine the rate of glyphosate dissipation from water under northern Manitoba conditions. Additionally, the major local mechanism for removal from the water column was to be identified. A series of six artificial pools was placed in staggered formation at the centre of the 1986 Manfor Ltd. experimental spray block. The spray block was located approximately 63 kilometres north of The Pas in north-western Manitoba. Three pools contained only water and three contained water and sediments. The pools were calibrated for volume and were each sampled seven times within a thirty-day period after herbicide application. Three sampling sites in a nearby natural pond were treated in a similar fashion and provided a comparison to the natural state. The study confirmed that the presence of sediments was essential for rapid removal of glyphosate from water in northern Manitoba. Glyphosate residues were found to dissipate rapidly from water when sediments were present; 50% of the initial concentration dissipated within approximately 22 hours of application. Maximum AMPA residues encountered in the water were 2.2% of the glyphosate applied. AMPA residues also dissipated rapidly from the water when sediments were present. AMPA dissipation to 1% of applied glyphosate was predicted within approximately 3.9 days of initial herbicide application. Glyphosate residues were comparatively persistent in sediments as evidence of dissipation was not detected during the thirty-day term of the study period. Beckett, R., D. M. Hotchin, and B. T. Hart. 1990. Use of field-flow fractionation to study pollutant-colloid interactions. Journal of Chromatography 517: 435-48. A new approach has been developed utilizing field-flow fractionation (FFF) methods to determine adsorption density distributions across the size spectrum of a particulate sample. The approach has been tested using sedimentation FFF to separate colloidal matter, concentrated from river water, to which radiolabelled pollutants (orthophosphate, glyphosate, atrazine) have been adsorbed. It is expected that the methodology will be expanded to cover the broad size range from 1 nm to 100 .m.m by using a number of FFF subtechniques (including sedimentation FFF): the use of various sensitive analytical methods will extend the range of adsorbates that can be studied. This provides a powerful new method for studying the pollutant-particle interactions occurring in environmental samples, as well as other similar systems. (Australia). Bowmer, K. H., P. M. D. Boulton, D. L. Short, and M. L. Higgins. 1986. Glyphosate-sediment interactions and phytotoxicity in turbid water. Pesticide Science 17: 79-88. See Plant and Soil Residues Section.

2.

3.

4.

5.

171

Water Quality 6. Chen, Y. L., H. C. Chiang, L. Q. Wu, and Y. S. Wang. 1989. Residues of glyphosate in an aquatic environment after control of water hyacinth (Eichhornia crassipes). Weed Research 34: 117-22. Glyphosate isopropylamine was surface sprayed at 1.166 and 1.02 kg/ha in model and field experiments, resp., for the control of E. crassipes in aquatic environments. Good control was obtained in ponds and irrigation canals. Residues dissipated rapidly with 1 ppm remaining immediately after treatment, 22% of the maximum concentration remaining in ponds 1 day after treatment and no residue detectable in canals after 4 hours. (Taiwan). Comes, R. D., V. F. Bruns, and A. D. Kelley. 1976. Residues and persistence of glyphosate in irrigation water. Weed Science 24: 47-50. Neither glyphosate (N-(phosphonomethyl)glycine) nor the soil metabolite aminomethylphosphonic acid were detected in the first flow of water through the two canals following application of glyphosate at 5.6 kg per ha to ditchbanks when the canals were dry. Soil samples collected the day before canals were filled (about 23 weeks after treatment) contained about 0.35 ppm glyphosate and 0.78 ppm aminomethylphosphonic acid in the 0 to 10-cm layer. When glyphosate was metered into the water at a rate calculated to provide 150 ppb in the canal water at a single site on two flowing canals, about 70% of the glyphosate was accounted for 1.6 km downstream from the application site. Thereafter, the rate of disappearance diminished, and about 58% of the applied glyphosate was present at the end of the canals 8 or 14.4 km downstream from the introduction sites. DeSnoo, G. R., and A Wegener-Sleeswijk. 1993. Use of pesticides along field margins and ditch banks in the Netherlands. Mededelingen Faculteit Landbouwkundige En Toegepaste Biologische Wetenschappen Universiteit Gent 58, no. (3A) : 921-26. (The Netherlands). Edwards, W. M., G. B. Triplett, and R. M. Kramer. 1980. A watershed study of glyphosate transport runoff. Journal of Environmental Quality 9: 661-65. Glyphosate (N-phosphonomethyl glycine) formulated as Roundup herbicide, was applied on 0.3- to 3.1-ha watersheds at rates of 1.10-, 3.36-, and 8.96-kg/ha as a preseeding herbicide in the notillage establishment of fescue and corn. Runoff from natural rainfall following early springtime treatment was measured and analyzed to define concentration and transport of glyphosate under these conditions. The highest concentration of glyphosate (5.2 mg/l) was found in runoff occurring 1 day after treatment at the highest rate. Glyphosate (2 g/l) was detected in runoff from this watershed up to 4 months after treatment. For the lowest application rates, maximum concentration of the herbicide in runoff was less than 100 g/l for events occurring 9-10 days after application and decreased to less than 2 g/l within 2 months of treatment. The maximum amount of glyphosate transport by runoff was 1.85% of the amount applied, most of which occurred during a single storm on the day after application. In each of the three study years, herbicide transport in the first runoff event following treatment accounted for 99% of the total runoff transport on one watershed. Glyphosate residues in the upper 2.5 cm of treated soil decreased logarithmically with the logarithm of time; they persisted several weeks longer than in the runoff water. Environment Canada. 1990. Pesticide research and monitoring annual report 1988-1989. Environment Canada , ISBN 0-662-18461-0; Publication no: DSS cat no En 40-11-13-1989E. This report summarizes the results of pesticide research and monitoring projects undertaken during 1988-89 in the Atlantic, Quebec, Ontario, Western and Northern, and Pacific and Yukon regions of Environment Canada as well as at the National Wildlife Research Centre. The projects undertaken included studies on the effects of herbicide spraying, the persistence of pesticides in subsurface drainage, the contamination of groundwater by pesticides, new techniques for monitoring pesticides in the environment, the toxicity of herbicide runoff on the growth and productivity of algae and aquatic invertebrates, the levels and effects of pesticides in birds, a compilation of the environmental fate and toxicity of specific pesticides, effects of pesticides on non-target organisms and the toxicity of glyphosate to juvenile salmon and rainbow trout. Guidelines for testing pesticide toxicity to non-target plants, data bases on the use of agricultural habitats by birds and an agricultural bird census were also developed. The report includes a decision document summarizing the environmental fate and effects of clofentizine, a miticide. 172

7.

8.

9.

10.

Water Quality 11. Faust, M., R. Altenburger, W. Boedeker, and L. H. Grimme. 1994. Algal toxicity of binary combinations of pesticides. Bulletin of Environmental Contamination and Toxicology 53 , no. 1:134-41 . See Aquatic Invertebrates and Algae Section. Feller, M. C. 1989. Effects of forest herbicide applications on streamwater chemistry in southwestern British Columbia. Water Resources Bulletin 25: 607-16. The herbicide glyphosate was applied to portions of two watersheds in southwestern British Columbia to kill vegetation that was competing with Pseudotsuga menziesii (Douglas-fir) plantations. This application had little significant effect on streamwater chemistry (K+, Na2+, Mg2+, Ca2+, Cl-, NO3-, NH4+, PO4-3, SO4=, and SiO2 concentrations, electrical conductivity, and pH) when vegetation cover in a watershed was reduced by 4%, but had significant (P<0.05) effects, which lasted for at least five years, when cover was reduced by 43%. In this case, most parameters increased in value following the application, with K+ and Mg2+ concentrations and pH values exhibiting the most prolonged increases and NO3- concentrations exhibiting the greatest percentage increases. Sulphate and dissolved SiO2 concentrations decreased following the application. Streamwater chemical fluxes showed similar trends to concentrations except that changes in fluxes were less significant and no decreases were observed. Forest management induced losses of NO3-N in streamwater during the first five posttreatment years in the study area decreased in the order: herbicide application (approximately 40 kg/ha) > clearcutting and slashburning (approximately 20 kg/ha) > clearcutting (approximately 10 kg/ha). In watersheds similar to those of the study area, herbicide application is likely to have a greater impact on streamwater chemistry, in general, than would clearcutting or clearcutting followed by slashburning. Feng, J. C., and D. G. Thompson. 1989. Persistence and dissipation of glyphosate in foliage and soils of a Canadian coastal forest watershed. Proceedings of the Carnation Creek Herbicide Workshop. B.C.: Forest Research Development Agreement, Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 65-87. See Plant and Soil Residues Section. Feng, J. C., D. G. Thompson, and P. E. Reynolds. 1989. Fate of glyphosate in a forest stream ecosystem. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada. FRDA Report 063, pp. 4564. Residues of glyphosate and its major metabolite aminomethyl phosphonic acid (AMPA) were monitored in stream water, bottom sediments and suspended sediments for a one year period following aerial application of glyphosate (2.0 kg/ha) to the Carnation Creek watershed of Vancouver Island, British Columbia. Analysis of deposit collector samples indicated an actual deposit of 1.88 kg/ha within the target zone and that less than 0.1% of the full rate impinged on surfaces beyond 8 m from the spray zone boundary. The highest concentrations of glyphosate (6.80 g/g dry mass) were observed in bottom sediments of a directly sprayed tributary, with low residue levels (less than 0.2 g/g) persisting in this substrate throughout the monitoring period. Suspended sediment samples collected from the main stream channel, indicated low levels of glyphosate input (0.10 g/L in conjunction with the first five storm events (23-66 days post application). The highest stream water residue observed (162 g/L) occurred in a directly sprayed tributary, 2 h post-application and decreased rapidly to 37g/L after 16 h. Transient increases in glyphosate concentrations were associated with first rainfall event (39 mm) which occurred 23 h post-application. Subsequent dissipation of residues in the two tributaries receiving direct glyphosate applications was rapid, such that no quantifiable residues were found after 96 h post-application. No quantifiable residues were found in tributaries buffered with a 10-m vegetation strip during the monitoring period. . 1990. Fate of glyphosate in a Canadian forest watershed. I. Aquatic residues and offtarget deposit assessment. Journal of Agricultural Food and Chemistry 38: 1110-1118. Glyphosate and AMPA residues in oversprayed and buffered streams were monitored following application of Roundup (2.0 kg/ha) to 45 ha of a coastal British Columbia watershed. Maximum glyphosate residues (stream water, 162 g/L; sediments, 6.80 g/g dry mass; suspended sediments, < 173

12.

13.

14.

15.

Water Quality 0.03 g/L) were observed in two intentionally oversprayed tributaries, dissipating to < 1 g/L within 96 h postapplication. Buffered streams were characterized by very low glyphosate residue levels (2.4-3.2 g/L in streamwater). Results of the off-target deposit assessment indicated < 0.1% of applied glyphosate at 8 m from the spray boundary. Increases in residue levels were observed in relation to the first storm event postapplication. Ratios of maximum stream water concentrations of glyphosate observed in buffered and oversprayed tributaries relative to literature toxicity values indicated a substantial margin of safety under either operational or worst case scenarios. 16. Gauch, R., U. Levenberger, and U. Mueller. 1989. Determination of glyphosate herbicide and its main metabolite aminomethylphosphonic acid AMPA in drinking water by HPLC. Zeitschrift Fuer Lebensmittel-Untersuchung Und -Forschung 188: 36-38. A method for the determination of glyphosate and its major metabolite aminomethylphosponic acid (AMPA) is described. With a detection limit of 0.02 m.g/l. The method suitably fulfils the requirements of the Swiss legislation (tolerance value of 0.1 m.g/l water). The compounds are derivatized directly in the original water sample with 9-fluorenylmethyl chloroformate (FMOCCL) in order to obtain extractable and fluorescent derivatives. These are extracted with organic solvents and determined by HPLC using a fluorescence detector. Neither of the compounds could be detected in 151 tap water samples from the Canton of Berne. (Germany). Gerritse, R. G., J. Beltran, and F. Hernandes. 1996. Adsorption of atrazine, simazine, and glyphosate in soils of the Gnangara Mound, Western Australia. Australian Journal of Soil Research 34, no. 4: 599-607. See Plant and Soil Residues Section. Goldsborough, L. G., and A. E. Beck. 1989. Rapid dissipation of glyphosate in small forest ponds. Archives of Environmental Contamination and Toxicology 18: 537-44. Glyphosate was applied to the water surface of four small boreal forest ponds and six in situ microcosms at a rate of 0.89 kg a.i./ha. Water samples collected over a period of up to 255 days were analyzed for glyphosate and its primary metabolite aminomethyl-phosphonic acid (AMPA). Glyphosate dissipated rapidly from all ponds with first order half-lives ranging from 1.5 to 3.5 days. The slowest dissipation rate occurred in the pond with the most calcareous water and sediments. Glyphosate remained at or above the treatment concentration in microcosms containing only water but decreased rapidly in the presence of sediments. AMPA levels in ponds and microcosms were consistently low. Concentrations on microcosm wall samples were temporally variable, probably a result of adsorption to periphytic biofilms. Glyphosate in the sediments of treated microcosms generally increased with time during the period of observation. These results confirm that glyphosate dissipates rapidly from the surface waters of lentic systems, and suggest that sediment adsorption or biodegradation were the major means of glyphosate loss from the water column. Goldsborough, L. G., and D. J. Brown. 1987. Effects of aerial spraying of forestry herbicides on aquatic ecosystems. Part III. Bioassay of the effect of glyphosate on carbon fixation by intact periphyton communities. Manitoba Environment and Workplace Safety and Health. Water Standards and Studies Report 873. See Aquatic Invertebrates and Algae Section. . 1993. Dissipation of glyphosate and aminomethylphosphonic acid in water and sediments of boreal forest ponds. Environmental Toxicology and Chemistry 12: 1139-47. Three small ponds in the boreal forest of southern Manitoba were treated with an aerial application of 2.1 kg/ha glyphosate. Two of the ponds had been treated the preceding year with 0.9 kg/ha glyphosate, whereas the third was previously untreated. Foliage samples of plants bordering the ponds were collected immediately after treatment, and water and sediment samples were collected over a period of 265 d after treatment. Samples were analyzed for glyphosate and its first metabolite, aminomethylphosphonic acid (AMPA). We found that glyphosate dissipated rapidly from the surface waters of all ponds (dissipation half-lives of 3.5-11.2 d). AMPA residues were detected in water samples during the first 14 d after treatment, suggesting that herbicide degradation was occurring in the water column. However, not all applied herbicide was accountable in residues in the water. 174

17.

18.

19.

20.

Water Quality Glyphosate and AMPA increased in sediment samples to day 36, suggesting that sediment adsorption was a major sink for the herbicide. Glyphosate dissipation from the water column was biphasic, with a rapid initial phase followed by a slower phase. We hypothesize this was due to herbicide sorption and partitioning between suspended particulate matter and bottom sediments. As was found after the first pond treatment, glyphosate dissipation was slower in chemically alkaline ponds than in a more dilute pond, suggesting that herbicide complexation with ions in solution may be a significant factor affecting the rate of dissipation in standing water. 21. Gutierrez, E., F. Arreguin, R. Huerto, and P. Saldana. 1994. Aquatic weed control. International Journal of Water Resource Development 10, no. 3: 291-312. See Plant and Soil Residues Section. Heitkamp, M. A., W. J. Adams, and L. E. Hallas. 1992. Glyphosate degradation by immobilized bacteria. Laboratory studies showing feasibility for glyphosate removal from waste water. Canadian Journal of Microbiology 38: 921-28. To evaluate immobilized bacteria technology for the removal of low levels of glyphosate (Nphosphonomethylglycine) from aqueous industrial effluents, microorganisms with glyphosate-degrading activity obtained from a fill and draw enrichment reactor inoculated with activated sludge were first exposed to glyphosate production wastes containing 500-2000 mg glyphosate/L. The microorganisms were then immobilized by adsorption onto a diatomaceous earth biocarrier contained in upflow Plexiglas columns. The columns were aerated, maintained at pH 7.0-8.0, incubated at 25 oC, supplemented with NH4NO3 (50 mg/L), and exposed to glyphosate process wastes dumped upflow through the biocarrier. Glyphosate degradation to aminomethylphosphonic acid was initially >96% for 21 days of operation at flows yielding hydraulic residence times (HRTs) as short as 42 min. Higher flow rates studies showed >98% removal of 50 mg glyphosate/L from the waste stream could be achieved at a HRT of 23 min. Glyphosate removal of >99% at a 37-min HRT was achieved under similar conditions with a column inoculated with a pure culture of Pseudomonas sp. strain LBr, a bacterium known to have high glyphosate-degrading activity. After acid shocking (pH 2.8 for 18 h) of a column of immobilized bacteria, glyphosate-degrading activity was regained within 4 days without reinoculation. Although microbial growth and glyphosate degradation were not maintained under low organic nutrient conditions in the laboratory, the low levels of degradable carbon (45-94 mg/L) in the industrial effluent were sufficient to support prolonged glyphosate-degrading activity. The results demonstrated that immobilized bacteria technology is effective in removing low levels of glyphosate in high-volume liquid waste streams. Hetherington, E. D. 1989. Carnation Creek floodplain hydrology, September 1984-September 1985. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, FRDA Report 063, pp. 27-44. The precipitation, streamflow, and groundwater regimes in the lower Carnation Creek floodplain are described for the 1-year period encompassing the application and monitoring of the environmental behaviour of the herbicide glyphosate. Rainfall just prior to the first herbicide application on the floodplain moistened the soil on well-drained areas and initiated ponding and flow of water in side channel depressions and swamps. The next rainfall on blocks containing monitored tributaries and seasonally flooded soil residue sampling sites came within 24 hours after herbicide application. Rainfall during the next 10 weeks was above average, providing ample opportunity for off-site movement and leaching of any mobile herbicide residues. However precipitation then remained below average for the rest of the study period. Surface water was continuously present in floodplain depressions and side channels until the following summer, with groundwater levels rising and falling in response to rainfall events. Surface soil layers on well-drained soil sampling sites would have remained moist but unsaturated and above the groundwater table throughout most of the study period. Of the three seasonally flooded soil sampling sites, two remained saturated and were frequently flooded until the following summer, while the other experienced unsaturated periods, particularly during the first 3 weeks after application, and only occasional flooding. The implications of these soil water regimes are that any mobile herbicide residues would have leached vertically downward on welldrained sites and under non-saturated soil conditions, but would have tended to move laterally from seasonally flooded stations during flooding and saturated conditions.

22.

23.

175

Water Quality 24. Judy, B. M., W. R. Lower, C. D. Miles, M. W. Thomas, and G. F. Krause. 1990. Chlorophyll fluorescence of a higher plant as an assay for toxicity assessment of soil and water. American Society for Testing and Materials. Symposium on use of Plants for Toxicity Assessment W. Wang, J. W. Gorsuch, and W. R. Lower (editors). Publ. 1091, pp. 308-18. Kataoka, H., K. Horii, and M. Makita. 1991. Determination of the herbicide glyphosate and its metabolite aminomethylphosphonic acid by gas chromatography with flame photometric detection. Agricultural Biology and Chemistry 55: 195-98. See Plant and Soil Residues Section. Lasater, B. M., and S. G. Berk. 1991. A toxicity assay for the marine environment based on INT reduction in marine ciliates. 91st General meeting of the American Society for Microbiology. p. 321. Legris, J., and G. Couture. 1989. Rsidus de glyphosate dans l'eau et les sdiments suite des pulvrisations terrestres en milieu forestier en 1986. Government of Quebec, Ministry of Energy and Resources, Environmental Studies Service. No. 3322 26 p. As a part of its plantation upkeep program, the Ministry of Energy and Resources has been using glyphosate (RoundupTM) since 1985 to control competing vegetation in reforested areas. In 1986, the Service des tudes environmentales carried out an environmental follow-up on glyphosate ground spraying in order to complete and confirm the results obtained in 1985. The present study was designed to provide data on the residual quantities of glyphosate in water and sediments in lotic and lentic environments affected by the spraying. A VoomjetTM cluster-nozzles sprayer mounted on a carrier was used to apply 1.5 kg a.i./ha. Samples were taken from eight streams draining the treated areas, each stream being protected by buffer zones 30-metres wide. In most cases, 3 consecutive water samples were collected at 12-hour intervals within a period of three days following treatment. The sampling sites were located at distances of 0, 0.1, 0.5 and 1 km from the area treated. None of the 60 water samples collected at these eight streams contained detectable levels of glyphosate (< 1.00 g/l). However, two positive results were detected in a drainage channel adjacent to a sprayed area (16.9 g/l) and in a water-course (3.16 g/l) draining an area treated with another spraying system. In both cases, there had been precipitation before sampling. Sediments were collected at the 0 km station in two of the streams studied 2 days and 1, 3, 7 and 10 weeks after spraying. Nine of the ten samples contained no detectable residues (< 0.05 g/g dry weight). The only positive values (0.100 g/g) was detected after 10 weeks. In a lentic environment, three ponds were directly exposed to spraying, and the maximum concentration found in the water was 2,808 g/l one hour after application. After 24 hours, the average concentration detected was only 288 g/l. In the sediments of those ponds, the residues rose from a maximum of 1.61 g/g after one week to an average value of 0.4 g/g after eleven weeks. A comparison between the residues detected in this study and the available toxicological data indicates that no measurable effects may be expected in the environments studied. . 1990. Rsidus de glyphosate dans un cosystme forestier suite des pulvrisations ariennes au Qubec en 1987. Ministry of Energy and Resources, Environmental Studies Services. ER90-3085. 35 p. See Plant and Soil Residues Section. Lveill, P., J. Legris, and G. Couture. 1993. Bilan des vrifications ponctuelles en milieu lotique la suite de pulvrisations de glyphosate en milieu forestier de 1989 1991. FQ93-3030. Gouvernement du Qubec, ministre des Forts, Direction de l'environment, Service du suivi environmental. 21 p. The ministre des Forts has used glyphosate (VisionTM) since 1985. This herbicide is used principally to control competing vegetation in softwood plantations. The Department's Service du suivi environmental is responsible for assessing the impact of glyphosate sprayings. For this purpose, the Service implements an environmental follow-up program and carries out spot checks which make it possible to observe and validate the product's pattern under the operational conditions of Quebec spraying program. This report presents the results of the spot checks carried out in lotic environments between 1989 and 1991. 176

25.

26.

27.

28.

29.

Water Quality Nine water samples were collected from eight watercourses. No residue was detected in the samples (<1.0 g/L). These results are in accord with previous data and can be explained by such factors as the product's relative lack of mobility (high degree of adsorption into soil particles), the effectiveness of buffer zones, and respect for other spraying standards. According to the results obtained for the 57 sediment samples collected from the 36 streams checked, 27 showed no detectable residue (<0.050 g/g). In the nine streams found to contain residues, concentrations ranged from 0.078 to 3.08 g/g. In cases of low sporadic contamination, the presence of the product might be due to contaminated soil particles flushed into the watercourse by heavy rains. In certain cases, the presence of glyphosate in sediments may be attributed to a failure to respect the buffer zone along the watercourse. According to this study, the low residue levels observed suggest that no appreciable toxic effect on aquatic organisms should be anticipated. 30. Mamarbachi, G., N. Nadeau, and M. Carmichael. 1989. "Mthodes d'analyse du glyphosate dans l'eau, l'air, le sol, les sdiments, la vgetation et les tissues animaux. " Government of Quebec, Ministry of Energy and Resources, Environmental studies. 19 p. Miles, C. J., L. R. Wallace, and H. A. Moye. 1986. Determination of glyphosate herbicide and (aminomethyl)phosphonic acid in natural waters by liquid chromatography using pre-column fluorogenic labelling with 9-fluorenylmethyl chloroformate. Journal of the Association of Analytical Chemistry 69: 458-61. The effect of glyphosate on nontarget organisms and its overall environmental fate cannot be fully evaluated unless techniques possessing suitable sensitivity and selectivity are available. An analytical method has been developed for determination of glyphosate herbicide and its major metabolite, (aminomethyl)phosphonic acid (AMPA), in natural waters. Sample pretreatment consisted of filtration, addition of phosphate buffer, concentration by rotary evaporation, and a final filtration before derivatization with 9-fluorenylmethyl chloroformate. The derivatives were separated by anion exchange liquid chromatography and measured with a fluorescence detector. Standard curves were linear over 3 orders of magnitude and minimal detectable quantities were 10 ng/ml for glyphosate and 5 ng/ml for AMPA. The 20-fold concentration factor realized in sample preparation corresponds to PPB method detection limits for glyphosate and AMPA in natural waters. Recovery and storage studies were performed and are discussed. Miller, J. J., N. Foroud, B. D. Hill, and C. W. Lindwall. 1992. Herbicides in surface runoff soil and ground-water under different agricultural management practices in southern Alberta. Canadian Journal Soil Science 72: 329. Miller, J. J., B. D. Hill, C. Chang, and C. W. Lindwall. 1995. Residue detections in soil and shallow groundwater after long-term herbicide applications in southern Alberta. Canadian Journal of Soil Science 75, no. 3: 349-56. See Plant and Soil Residues Section. Neary, D. G., and J. L. Michael. 1995. The role of herbicides in protecting long-term sustainability and water quality in forest ecosystems. Second International Conference on Forest Vegetation Management. Rotorua, New Zealand. R.E. Gaskin and J.A. Zabkiewicz (compilers). FRI Bulletin 192, pp. 161-63. The use of herbicides for controlling competing vegetation during stand establishment can be beneficial to forest ecosystem sustainability and water quality by minimizing off-site soil loss. In addition, the organic residues of forestry herbicides do not adversely impair water quality. Newton, M., L. M. Horner, J. E. Cowell, D. E. White, and E. C. Cole. 1994. Dissipation of glyphosate and aminomethylphosphonic acid in North American forests. Journal of Agricultural and Food Chemistry 42, no. 8: 1795-802. See Plant and Soil Residues Section.

31.

32.

33.

34.

35.

177

Water Quality 36. Newton, M., K. M. Howard, B. R. Kelpsas, R. Danhaus, C. M. Lottman, and S. Dubelman. 1984. Fate of glyphosate in an Oregon USA forest ecosystem. Journal of Agricultural and Food Chemistry 32: 1144-51. See Mammals Section. Ogner, G. 1985. Water quality following forest fertilization. Part 3. Long-term effects of fertilizer application and effects of spraying glyphosate on the Hegga river, Norway. 5. Norsk Institutt for Skogforskning. 10 p. (Norway). . 1987. Glyphosate application in forest-ecological aspects. II. The quality of water leached from forest soil lysimeters. Scandinavian Journal of Forest Research 2: 469-80. The vegetation of four different forest soils in lysimeters was killed with glyphosate. New vegetation became established in some of the treated lysimeters the year after application. During this year, leachates from lysimeters without vegetation had nitrate concentrations up to 4.5 mM NO3- in the fall. A corresponding increase in leached cations was evident. The settlement of one raspberry or one willow herb in treated lysimeters effectively hindered nitrate leaching as soon as the plants were sufficiently large (in mid-summer). As plants became established in treated lysimeters, the chemistry of these leachates approached those of untreated lysimeters. The absolute effect of even small plants in controlling water regime and water chemistry was evident. (Norway). . 1987. Glyphosate application in forest-ecological aspects. IV. The water quality of forest brooks after routine application. Scandinavian Journal of Forest Research 2: 499-508. The water quality of forest brooks was monitored after glyphosate application to two Norwegian forest areas. The application was done according to common forestry practice. The major effect on water quality of glyphosate is to increase the concentration of NO3-. The maximum concentration found was 80-90 M NO3- the first and second winters after application. Without glyphosate application, the level was estimated to be 20-40 M lower. Generally the increase in NO3- was smaller or undetectable. Only at the higher levels of NO3- was an increase in water acidity observed. This occurred the first summer after application and was in the range 1-3 M H+. No other effects on the water quality resulting from glyphosate application were detected. Glyphosate application as part of common forestry management in Norway gives only small changes in the water quality of forest brooks. As long as small areas of the forest are treated the consequences of the changes in water quality appear nonsignificant. (Norway). . 1987. Glyphosate application in forest-ecological aspects. V. The water quality of forest brooks after manual clearing or extreme glyphosate application. Scandinavian Journal of Forest Research 2: 509-16. The water quality of forest brooks was monitored after manual clearing of forest plantations and in an area treated with glyphosate for two consecutive years. In both cases an increase in the concentration of NO3- was evident. Manual clearing gave concentrations of 20-60 M NO3- during winter and less than 10 M during the growth season, which is in the same range as usually found in glyphosate-treated forests. The doubly-treated glyphosate area showed increased leaching of NO3 during summer, given 30-60 M NO3 as the maximum value the first summer after the second application. The summer leaching decreased the following two years, giving brook water with 5-25 M NO3 . No effect of the treatments was found for the other nutrients measured. Both manual clearing and glyphosate application in forest plantations give the same small effect on brook water quality. The effects appear not to be of ecological importance. (Norway). Paveglio, F. L., K. M. Kilbridge, C. E. Grue, C. A. Simenstad, and K. L. Fresh. 1996. Use of rodeo and X-77 spreader to control smooth cordgrass (Spartina alterniflora) in a southwestern Washington Estuary: 1. Environmental fate. Environmental Toxicology and Chemistry 15, no. 6: 961-68. A 1-ha plot with smooth cordgrass (Spartina alterniflora) at three locations in Willapa Bay, a southwestern Washington estuary, was aerially treated with 4.7 L/ha Rodeo and 0.9 L/ha X-77 Spreader to determine the fate of the herbicide formulation. Rates of spray deposit on filter pads placed over 178

37.

38.

39.

40.

41.

Water Quality treated intertidal mudflats did not differ among locations for glyphosate and aminomethylphosphonic acid (AMPA); however, deposit rates for nonylphenol polyethoxylates (NPEO) did differ among locations. Glyphosate concentrations in sediment cores from treated mudflats declined 51 to 72% during 119 days posttreatment (DPT); NPEO concentrations in sediment declined (42%) between spray day and 14 DPT. The highest concentrations of glyphosate, AMPA, and NPEO in seawater were found in off-site samples collected from the leading edge of the first high tide after application. Glyphosate and AMPA concentrations in depth-integrated seawater samples declined 73 and 42%, respectively, between the first high tide immediately following application and the second high tide at 1 DPT. Glyphosate concentrations in Spartina stems from treated plots declined 91 to 99% between I and 28 DPT; whereas, AMPA declined (86 and gt 96%) during this period. Comparison of maximum concentrations for glyphosate in seawater from this study with acute toxicity values in the literature indicates that under worst-case conditions direct effects to aquatic organisms would not be likely. 42. Pawlizki, K. H. 1987. Effects of herbicides used in orchards and vineyards on man and the environment. Gesude Pflanzen 39: 486-96. See Plant and Soil Residues Section. Petite, V., R. Cabridenc, R. P. J. Swannell, and R. S. Sokhi. 1995. Review of strategies for modelling the environmental fate of pesticides discharged into riverine systems. Environment International 21, no. 2: 167-76. Pesticides are often produced and stored in large quantities near rivers posing a potential hazard for the aquatic environment. Accidental incidents such as storage facility fires are of major concern as significant amounts of pesticide chemicals can enter the nearby riverine system, possibly causing considerable environmental damage. This paper discusses and reviews the major physical, chemical, and microbiological fate processes of selected herbicides in riverine systems. Glyphosate, paraquat, and diquat herbicides have been selected for discussion as they are widely used and because they degrade in freshwater mainly by well-defined fate processes. The paper concentrates on biodegradation, sorption, and photolysis, the primary fate processes by which these herbicides degrade. Strategies for mathematically modelling the environmental fate of pesticides in rivers are reviewed and areas of future work identified. Piccolo, A., and G. Celano. 1994. Hydrogen-bonding interactions between the herbicide glyphosate and water-soluble humic substances. Environmental Toxicology and Chemistry 13, no. 11: 1737-41. In this study, infrared spectroscopy (IR) was used to investigate the mechanism of interaction between the widely used herbicide glyphosate [N-(phosphonomethyl) glycine] and a water-soluble humic acid (HA) obtained from a purified Leonardite. Aliquots of the purified humic solution were set to pH 4, 5, and 6, and added with glyphosate volumes corresponding to increasing percentages of total acidity of the water-soluble HA. The IR spectra of the HA-glyphosate complexes showed two bands at 1,168 and 1,090/cm for the stretching of the P=O and P-O bonds, respectively, as they appear in the spectrum of glyphosate alone where the phosphono group is involved in hydrogen bondings occurring in glyphosate dimers. Conversely, when the complex solutions were brought back to the starting pH levels (4, 5, and 6), the two IR bands of the phosphono group were shifted to higher frequencies (1,195 and 1,134/cm, respectively), suggesting that the P-O bonds were no longer involved in hydrogen bondings. This behaviour supports the occurrence of hydrogen bondings in the glyphosate-HA complex. In fact, the IR band shift in the herbicide-HA complex indicates that, because the glyphosate molecules are distributed along the humic macromolecule, they are too far apart to form the glyphosate dimers again when the initial conditions are restored. The hydrogen-bonding interactions between HA and glyphosate evidenced in this study may well be responsible for the transport of glyphosate in lower soil depths through the adsorption of glyphosate on water-soluble humic substances. (Italy). Reinert, K. H., and J. H. Rodgers. 1987. Fate and persistence of aquatic herbicides. Reviews of Environmental Contamination and Toxicology 98: 61-98. The aquatic fate and persistence of acrolein, amitrole, copper sulfate and complexes, dalapon, dicamba, dichlobenil, 2,4-D, diquat, endothal, fenac (chlorfenac), fluridone, glyphosate, simazine and

43.

44.

45.

179

Water Quality xylenes, all registered for use in the USA, are discussed. Data obtained from a model developed to estimate herbicidal behaviour in aquatic systems are included. 46. Reynolds, P. E., J. C. Scrivener, L. B. Holtby, and P. D. Kingsbury. 1989. An overview of Carnation Creek herbicide study: historical perspective, experimental protocols, and spray operations. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, FRDA Report 063, pp. 1526. Research objectives, proposals and protocols for assessing the environmental impacts of applying Roundup (glyphosate) as a control of forest brush species at Carnation Creek were agreed to in 1984 by the British Columbia (BC) Ministry of Forests, BC Ministry of Environment and Parks, Canadian Forestry Service, Fisheries and Oceans Canada, and the Council of Forest Industries. Carnation Creek is a 10 sq km watershed in the high rainfall cedar-hemlock zone of coastal British Columbia, where the impacts of forestry practices on salmonids have been studied since 1970 under the Carnation Creek Experimental Watershed Project. Project objectives were to improve understanding of any impacts and the relative magnitude of clearcut logging and post-logging forestry practices on salmon and trout populations and their stream habitat. The Forest Pest Management Institute coordinated and administered the herbicide study as a component to the overall Carnation Creek Experimental Watershed Project. Roundup was applied aerially at 2.0 kg/ha to 41.7 ha of the watershed in September 1984. A 10-m pesticide free zone was maintained along the stream, but two tributary swamps were oversprayed as part of the study design. Short-term direct impacts to stream water, vegetation, soils and stream biota were assessed. Long-term indirect impacts on water quality, erosion processes, and stream biota were studied until at least June 1986. . 1989. A summary of Carnation Creek herbicide study results. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, B.C. FRDA Report 063, pp. 322-34. See Fish Section. . 1993. Review and synthesis of Carnation Creek Herbicide research. The Forestry Chronicle 69: 323-30. Carnation Creek Watershed (48o54'N,125o01'W), located on the west coast of Vancouver Island, was aerially treated with glyphosate (ROUNDUP) in September 1984. Various chemical and biological studies were conducted for up to 3 years after the treatment in order to study the environmental fate and impacts of the herbicide in a temperate coastal rain forest. In tributaries oversprayed with the herbicide, impacts on coho salmon (Oncorhynchus kisutch) and other aquatic organisms were short-term, and the impacts were considered to be acceptable. Residue movements within the watershed and residue inputs into the aquatic ecosystem were monitored in relation to autumn and winter storms. Glyphosate residues rapidly dissipated and degraded in the natural environment. After 1 year, remaining residues were strongly adsorbed to organic matter, soil particles, and/or stream bottom sediments, where they appeared to be inactivated and immobilized. Roy, D. N., S. K. Konar, S. Banerjee, D. A. Charles, D. G. Thompson, and R. Prasad. 1989. Persistence movement and degradation of glyphosate in selected Canadian boreal forest soils. Journal of Agricultural and Food Chemistry 37: 437-40. See Plant and Soil Residues Section. Rueppel, M. L., B. B. Brightwell, M. Schaefer, and J. T. Marvel. 1977. Metabolism and degradation of glyphosate in soil and water. Journal of Agricultural and Food Chemistry 25: 517-28. See Plant and Soil Residues Section.

47.

48.

49.

50.

180

Water Quality 51. Sacher, R. M. 1978. Safety of Roundup in the aquatic environment. Proceedings of the 5th International Conference on Aquatic Weeds. pp. 315-21 Wageningen, NL: European Weed Res. Soc. Roundup, a new broad spectrum herbicide, is a formulation of isopropylamine salt of N(phosphonomethyl)glycine. The product, which is highly effective in the control of both annual and perennial herbaceous plants, shows extremely favourable toxicology. Furthermore, the chemical is degraded by water and/or soil components. Detailed 14C-studies show that up to 50% of the carbon atoms within the compounds evolve as CO2 within 28 days. Analysis of the soil bound residues confirm that the major metabolites of the compound are aminomethylphosphonic acid and natural products. On two of three soils, the compound was 90% degraded in less than 12 weeks. Results are also presented on the rapid dissipation of the compound in ponds and irrigation canals, its negligible bioaccumulation and its safety to mammals, birds, fish and other aquatic organisms. Scrivener, J. C. 1989. Comparative changes in concentration of dissolved ions in the stream following logging, slash burning, and herbicide application (glyphosate) at Carnation Creek, British Columbia. Proceedings of the Carnation Creek Herbicide Workshop. Forest Resource Development Agreement, B.C. Ministry of Forests, Forestry Canada, FRDA Report, 063, pp. 197211. The patterns of dissolved ions were studied in the main channel, in two pristine tributaries, and in two tributary swamps (1600-trib, 2600-trib) for 14 yr prior to and for 2 yr after aerial applications of the herbicide ROUNDUP (glyphosate). Four percent (41.7 ha) of the main watershed and half of the watershed including the stream channel of tributary 1600 were sprayed with the herbicide. In all the streams electrical conductivity and the concentration of dissolved nitrate-N were negatively correlated with stream flow. After herbicide treatment these values were unchanged at minimum to median stream flow but conductivity increased 21% and nitrate concentrations increased 3-fold during periods of freshet. Increases that were 4 times as large had been observed when these watersheds were logged and the logging slash was burned. Phosphate-P concentration was not correlated with stream flow and it increased 2-fold after herbicide application. No changes in dissolved phosphate concentration had been observed after logging and slash burning. Speth, T. F. 1993. Glyphosate removal from drinking water. Journal of Environmental Engineering 119, no. 6: 1139-57. Activated-carbon, oxidation, conventional-treatment, filtration, and membrane studies are conducted to determine which process is best suited to remove the herbicide glyphosate from potable water. Both bench-scale and pilot-scale studies are completed. Computer models are used to evaluate the results. The activated-carbon results show that glyphosate adsorbs very strongly in distilled water, but has a much lower capacity in Ohio River water. The jar-test studies with an alum coagulant show that as turbidity is removed, so is glyphosate. The majority of the glyphosate removal occurs as turbidity is reduced below 2 nephelometric turbidity units (NTUs). Powdered-activatedcarbon treatment is ineffective. Ultrafiltration membranes and 0.45 mu m filters do not remove glyphosate in Ohio River water even though the effluent turbidity is reduced below 0.2 NTU. The oxidation results indicate that glyphosate is easily destroyed by chlorine and ozone. Chlorine dioxide, permanganate, and hydrogen peroxide are less successful. Subramaniam, V., and P. E. Hoggard. 1988. Metal complexes of glyphosate. Journal of Agricultural and Food Chemistry 36: 1326-29. Insoluble complexes of glyphosate with iron (III), copper(II), calcium, and magnesium ions are formed at near-neutral pH. This suggests a mechanism for the inactivation of glyphosate in contaminated groundwater. Thelen, K. D., E. P. Jackson, and D. Penner. 1995. The basis for the hard-water antagonism of glyphosate activity. Weed Science 43, no. 4: 541-48. Hard-water cations, such as Ca+2 and Mg+2, present in the spray solution can greatly reduce the efficacy of glyphosate. These cations potentially compete with the isopropylamine in the formulation for association with the glyphosate anion. 14C-Glyphosate absorption by sunflower was reduced in the presence of Ca+2. The addition of ammonium sulfate overcame the observed decrease 181

52.

53.

54.

55.

Water Quality in 14C-glyphosate absorption. Nuclear Magnetic Resonance (NMR) was used to study the chemical effects of calcium and calcium plus ammonium sulfate (AMS) on the glyphosate molecule. Data indicate an association of calcium with both the carboxyl and phosphonate groups on the glyphosate molecule. Initially, a random association of the compounds occurred; however, the reaction progressed to yield a more structured, chelate type complex over time. NH-4+ from AMS effectively competed with calcium for complexation sites on the glyphosate molecule. Data suggest that the observed calcium antagonism of glyphosate and AMS reversal of the antagonism are chemically based. 56. Tooby, T. E. 1985. Fate and biological consequences of glyphosate in the aquatic environment. In: The Herbicide Glyphosate. E. Grossbard, and D. Atkinson, pp. 206-17. London: Butterworths. Glyphosate, when used as recommended by the manufacturer, is unlikely to enter watercourses through run-off or leaching following terrestrial application. Clearly, significant residues could enter water when Roundup is used as an aquatic herbicide. In static water, residues decline fairly rapidly due mainly to adsorption onto particulate matter. In some flowing water the conditions could be such that glyphosate could remain in solution or adsorbed onto particulate matter in suspension, for some distance downstream from the point of application. Although the rate of degradation of residues depends on pH, temperature and the presence of microorganisms, most aquatic environments likely to receive Roundup treatment will provide the most ideal conditions for degradation. The field experience seems to support the laboratory predictions that residues will be adsorbed and will eventually break down. Therefore, it is unlikely that glyphosate will affect aquatic organisms at the concentrations found in the environment after use at the recommended rates. It is also unlikely that residues will be accumulated in fish tissues. Trotter, D. M., M. P. Wong, and R. A. Kent. 1990. Canadian water quality guidelines for glyphosate., Water Quality Branch, Inland Waters Directorate, Environment Canada, Ottawa, Ontario. 27 p. A literature review was conducted on the uses, fate, and effects of glyphosate on raw water for drinking water supply, freshwater aquatic life, agricultural water uses, recreational water quality and aesthetics, and industrial water supplies. The information is summarized in this publication. From it, water quality guidelines for the protection of specific water uses are recommended. Wan, M. T. K. 1986. The persistence of glyphosate and its metabolite amino-methyl-phosphonic acid in some coastal British Columbia streams. Environment Canada, Environmental Protection Service, Regional program, Pacific and Yukon Region. Report 85-01. From 1982 to 1984, the Environmental Protection Service, Pacific Region, conducted studies at the UBC Research Forest and Chilliwack/Sechelt Forest Districts to monitor the residues of glyphosate and its metabolite, amino-methyl-phosphonic acid (AMPA), in stream water and sediment. The objective of these studies was to evaluate the persistence of the two chemicals in coastal British Columbia streams protected and unprotected by a buffer zone during aerial and ground based operational sprays. Glyphosate residues in unprotected stream water reached a level of 0.023 mg/l within 2-3 hours following aerial application at 3 kg a.i./ha. These residues peaked at 0.100 mg/l following the first major rainstorm eight days later, and then dropped to levels below the detection limit of 0.005 mg/l within two weeks. In two protected streams, the highest water level detected was 0.025 mg/l, following several heavy rainstorms. In both the protected and unprotected streams, the concentration of AMPA in water was below the detection limit of 0.005 mg/l during the study period. One week after spray application, glyphosate residue in a sediment sample from the unprotected stream was 0.100 mg/kg. Peak levels of 0.400 mg/kg glyphosate were detected at three weeks and three months post-spray, and thereafter concentrations declined to 0.040 mg/kg which persisted for 574 days post-spray. AMPA residues were also found in sediments, peaking at 0.400 mg/kg three months after treatment. This metabolite showed a more gradual decline to 0.090 mg/kg which also persisted for the duration of the study (574 days). Glyphosate residue in a sediment sample from a protected stream was 0.200 mg/kg one week after an aerial spray application at 3 kg a.i./ha. Herbicide residues from vegetation along the stream bank or in the treated area may have washed into the creek following major rainstorm events. AMPA residues were detected only once in this stream at 0.100 mg/kg, 15

57.

58.

182

Water Quality days after treatment. Both glyphosate and AMPA residues were below the level of detection in protected stream sediments for the remainder of the study (600 days post-spray). 59. Wang, Y. S., J. H. Yen, Y. N. Hsieh, and Y. L. Chen. 1994. Dissipation of 2,4-D, glyphosate and paraquat in river water. Water, Air and Soil Pollution 72 : 1-7. Dissipation of herbicides in river water was determined by adding different concentrations of 2,4-D, glyphosate and paraquat to samples of river water. A small variation of dissipation of radioactivity for 14C-2,4-D in higher and lower concentrations and in different samples of river water was found. But about half the radioactivity disappeared from water samples of original glyphosate concentration at 100 mg/L and, in the case of 100 mu-g/L, only 11 to 22% remained in the samples of river water after 56 d incubation except the sample from Hsin-Tien River. More than 80% of paraquat remained in water samples. Determination of octanol-water partition coefficient (Kow) showed a large difference in amounts of 2,4-D partitioned in water phase at different pH values, 97.4% at the higher pH of ionic state and 5.2% at the lower pH of molecular state, implying that pH value of water might affect the bioaccumulation process of 2,4-D. The result showed that 95.0% of glyphosate present in water phase in ionic form (higher pH) and 82.3% in molecular form (lower pH), indicating that glyphosate might have no effect on the biomagnification, since most of glyphosate could be excreted with water by organisms. (Taiwan). Wang, Y., C. Jaw, and Y. Chen. 1994. Accumulation of 2,4-D and glyphosate in fish and water hyacinth. Water, Air and Soil Pollution 74 , no. 3-4 : 397-403 . See Fish Section. Wigfield, Y. Y., and M. Lanouette. 1990. Simplified liquid chromatographic determination of glyphosate and metabolite residues in environmental water using post-column fluorogenic labelling. Analytica Chimica Acta 233: 311-14. Glyphosate and its metabolite aminomethylphosphonic acid in environmental water were preconcentrated with an anion-exchange column, eluted with potassium citrate solution and determined directly by liquid chromatography with a post-column reactor and a fluorescence detector. The limit of detection and av. recovery were 1 microg/litre and 89.3% for glyphosate and 0.4 microg/litre and 86.3% for the metabolite. Wimberely, D. N., and S. G. Berk. 1991. Microplate analysis of dehydrogenase activity in protozoa a rapid aquatic toxicity assay. General Meeting of the American Society for Microbiology , 321. Zaranyika, M. F., and M. G. Nyandoro. 1993. Degradation of glyphosate in the aquatic environment: An enzymatic kinetic model that takes into account microbial degradation of both free and colloidal (or sediment) particle adsorbed glyphosate. Journal of Agricultural and Food Chemistry 41, no. 5: 838-42. The kinetics of the degradation of the herbicide glyphosate in distilled water and river water containing river sediment were investigated over a period of 72 days. No appreciable degradation of glyphosate was observed in distilled water, while rapid degradation occurred in the river water plus sediment from the outset, suggesting that the degradation is mainly microbial. An immediate 35% loss from solution of glyphosate due to adsorption to suspended sediment particles and deposition to the bottom sediment was observed in the river water plus sediment experiment. Subsequently, two linear rates of degradation were observed in the water phase of this experiment: an initial rapid degradation followed by a slower breakdown. An enzymatic kinetic model is presented showing that the rate of degradation of glyphosate (G) is given by -d (Delta-G)/dt = k-2 (G-B) + k-6 (GC-B), where k-6 and k-2 are the rate constants for sediment or colloidal particle absorbed glyphosate (GC) and the unadsorbed glyphosate (G), respectively, and the subscript B denotes microflora-bound. (Zimbabwe).

60.

61.

62.

63.

183

Das könnte Ihnen auch gefallen