TITLE: Allelopathy: The Effects of Chemicals Produced by
 Plants
 PUBLICATION DATE:  September 1994
 ENTRY DATE:  April 1995
 EXPIRATION DATE:  
 UPDATE FREQUENCY: 
 CONTACT:  Jane Gates
           Alternative Farming Systems Information Center
           National Agricultural Library
           Room 304, 10301 Baltimore Ave.
           Beltsville, MD  20705-2351
           Telephone:  (301) 504-6559
           FAX:  (301) 504-6409
           
 DOCUMENT TYPE:  text
 DOCUMENT SIZE:  340k (159 pages)
 
 
 ==============================================================
                                              ISSN:  1052-5378
 United States Department of Agriculture
 National Agricultural Library
 10301 Baltimore Blvd.
 Beltsville, Maryland  20705-2351
 
 Allelopathy:  The Effects of Chemicals Produced by Plants
 January 1990 - March 1994
 
 
 
 
 QB 94-56
 Quick Bibliography SeriesBibliographies in the Quick Bibliography Series of the
 National Agricultural Library, are intended primarily for
 current awareness, and as the title of the series implies, are
 not indepth exhaustive bibliographies on any given subject. 
 However, the citations are a substantial resource for recent
 investigations on a given topic.  They also serve the purpose
 of bringing the literature of agriculture to the interested
 user who, in many cases, could not access it by any other
 means.  The bibliographies are derived from computerized on-
 line searches of the AGRICOLA data base.  Timeliness of topic
 and evidence of extensive interest are the selection criteria.
 
 The author/searcher determines the purpose, length, and search
 strategy of the Quick Bibliography.  Information regarding
 these is available upon request from the author/searcher.
 
 Copies of this bibliography may be made or used for
 distribution without prior approval.  The inclusion or
 omission of a particular publication or citation may not be
 construed as endorsement or disapproval.
 
 To request a copy of a bibliography in this series, send the
 series title, series number and self-addressed gummed label
 to:
 
 U.S. Department of Agriculture
 National Agricultural Library
 Public Services Division, Room 111
 Beltsville, Maryland 20705-2351
 
 Allelopathy:  The Effects of Chemicals Produced by Plants
 January 1990 - March 1994
 
 
 
 Quick Bibliography Series:  QB 94-56
 Updates QB 92-50
 
 
 244 citations from AGRICOLA
 
 Henry Gilbert
 Reference and User Services Branch
 
 
 
 
 
 
 
 September 1994
 National Agricultural Library Cataloging Record:
 
 Gilbert, Henry
   Allelopathy : the effects of chemicals produced by plants: 
 1990 - March 1994.
   (Quick bibliography series ; 94-56)
   1. Allelopathy--Bibliography. I. Title.
 aZ5071.N3 no.94-56
 
 
 
 The United States Department of Agriculture (USDA) prohibits
 discrimination in its programs on the basis of race, color,
 national origin, sex, religion, age, disability, political
 beliefs, and marital or familial status.  (Not all prohibited
 bases apply to all programs).  Persons with disabilities who
 require alternative means for communication of program
 information (braille, large print, audiotape, etc.) should
 contact the USDA Office of Communications at (202) 720-5881
 (voice) or (202) 720-7808 (TDD).
 
 To file a complaint, write the Secretary of Agriculture, U.S.
 Department of Agriculture, Washington, D.C.  20250, or call
 (202) 720-7327 (voice) or (202) 720-1127 (TDD).  USDA is an
 equal employment opportunity employer.
 
 AGRICOLA
 
 Citations in this bibliography were entered in the AGRICOLA
 database between January 1979 and the present.
 
 SAMPLE CITATIONS
 
 Citations in this bibliography are from the National
 Agricultural Library's AGRICOLA database.  An explanation of
 sample journal article, book, and audiovisual citations
 appears below.
 
 JOURNAL ARTICLE:
 
   Citation #                                     NAL Call No.
   Article title.
   Author.  Place of publication:  Publisher.  Journal Title.
   Date.  Volume (Issue).  Pages.  (NAL Call Number).
 
 Example:
   1                             NAL Call No.:  DNAL 389.8.SCH6
   Morrison, S.B.  Denver, Colo.:  American School Food Service
   Association.  School foodservice journal.  Sept 1987. v. 41
   (8). p.48-50. ill.
 
 BOOK:
 
   Citation #                                   NAL Call Number
   Title.
   Author.  Place of publication:  Publisher, date. Information
   on pagination, indices, or bibliographies.
 
 Example:
   1                        NAL Call No.:  DNAL RM218.K36 1987
   Exploring careers in dietetics and nutrition.
   Kane, June Kozak.  New York:  Rosen Pub. Group, 1987.
   Includes index.  xii, 133 p.: ill.; 22 cm.  Bibliography:
   p. 126.
 
 AUDIOVISUAL:
 
   Citation #                                  NAL Call Number
   Title.
   Author.  Place of publication:  Publisher, date.
   Supplemental information such as funding.  Media format
   (i.e., videocassette):  Description (sound, color, size).
 
 Example:
   1                    NAL Call No.: DNAL FNCTX364.A425 F&N AV
   All aboard the nutri-train.
   Mayo, Cynthia.  Richmond, Va.:  Richmond Public Schools,
   1981.  NET funded.  Activity packet prepared by Cynthia
   Mayo.  1 videocassette (30 min.): sd., col.; 3/4 in. +
   activity packet.
 Allelopathy:  The Effects of Chemicals Produced by Plants
 
 SEARCH STRATEGY
 
      SET   DESCRIPTION
 
       S1    ALLELOPATHY?/TI,DE
 
       S2    ALLELOCHEM?/TI,DE
 
       S3    ALLELOPATH?/TI,DE OR ALLELOCHEM?/TI,DE
 
       S4    S3 AND S4
 
 
 Allelopathy:  The Effects of Chemicals Produced by Plants
                                
      
 1                                   NAL Call. No.: 450 P5622
 (20S)-4 alpha-methyl-24-methylenecholest-7-en-3 beta-ol, an
 allelopathic sterol from Typha latifolia.
 Della Greca, M.; Mangoni, L.; Molinaro, A.; Monaco, P.;
 Previtera, L. Oxford : Pergamon Press; 1990.
 Phytochemistry v. 29 (6): p. 1797-1798; 1990.  Includes
 references.
 
 Language:  English
 
 Descriptors: Typha latifolia; Anabaena flos-aquae; Chlorella
 vulgaris; Aquatic weeds; Weed control; Allelopathy; Chemical
 constituents of plants; Sterols
 
 
 2                                  NAL Call. No.: QD415.A1J6
 2,2'-oxo-1,1'-azobenzene: a microbially transformed
 allelochemical from 2,3-benzoxazolinone. I.
 Nair, M.G.; Whitenack, C.J.; Putnam, A.R.
 New York, N.Y. : Plenum Press; 1990 Feb.
 Journal of chemical ecology v. 16 (2): p. 353-364; 1990 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Soil analysis; Pesticide residues; Metabolites;
 Herbicides; Azo compounds; Allelopathy; Secale cereale
 
 Abstract:  2,2'-Oxo- 1,1'-azobenzene (AZOB), a compound with
 strong herbicidal activity, was isolated and characterized
 from a soil supplemented with 2,3-benzoxazolinone (BOA). A
 parallel experiment with
 6-methoxy-2,3-benzoxazolinone (MBOA) yielded AZOB as well as
 its mono- (MAZOB) and dimethoxy- (DIMAZOB) derivatives. These
 compounds were produced only in the presence of soil
 microorganisms, via possible intermediates, I and II, which
 may dimerize or react with the parent molecule to form the
 final products. In the case of MBOA, it was shown that
 demethoxylation precedes the oxidation step. Although BOA and
 2,4-dihydroxy-1,4(2H)-benzoxazin-3-one (DIBOA) were leached
 out of rye residues, there were no detectable amounts of the
 biotransformation products in the soil. When BOA was mixed
 with soil and rye residue, either under field conditions or in
 vitro, AZOB was detected. Levels of free BOA in the soil were
 greatly reduced by incubation with rye residue. AZOB was more
 toxic to curly cress (Lepidium sativum L.) and barnyardgrass
 (Echinochloa crusgalli L.) than either DIBOA or BOA.
 
 
 3                                  NAL Call. No.: QD415.A1J6
 2,2'-oxo-1,1'-azobenzene: microbial transformation of rye
 (Secale cereale L.) allelochemical in field soils by
 Acinetobacter calcoaceticus. III. Chase, W.R.; Nair, M.G.;
 Putnam, A.R.; Mishra, S.K.
 New York, N.Y. : Plenum Press; 1991 Aug.
 Journal of chemical ecology v. 17 (8): p. 1575-1584; 1991 Aug. 
 Includes references.
 
 Language:  English
 
 Descriptors: Secale cereale; Plant composition;
 Allelochemicals; Acinetobacter calcoaceticus; Allelopathy;
 Microbial activities
 
 Abstract:  Acinetobacter calcoaceticus, a gram-negative
 bacterium isolated from field soil, was found to be
 responsible for the biotransformation of 2(3H)-benzoxazolinone
 (BOA) to 2,2'-oxo-1,1'-azobenzene (AZOB). Experiments were
 conducted to evaluate the transformation of BOA to AZOB by
 this microbe in sterile and nonsterile soil. Transformation
 studies with soils inoculated with A. calcoaceticus indicated
 that the production of AZOB increased linearly with the
 concentration of BOA in sterile soil and showed a quadratic
 trend in nonsterile soils. This also indicated that all soil
 types studied for the transformation experiments might contain
 A. calcoaceticus capable of the conversion of
 benzoxazolinones.
 
 
 4                                  NAL Call. No.: QD415.A1J6
 2,2'-oxo-1,1'-azobenzene: selective toxicity of rye (Secale
 cereale L.) allelochemicals to weed and crop species. II.
 Chase, W.R.; Nair, M.G.; Putnam, A.R.
 New York, N.Y. : Plenum Press; 1991 Jan.
 Journal of chemical ecology v. 17 (1): p. 9-19; 1991 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Secale cereale; Plant composition;
 Allelochemicals; Toxicity; Bioassays; Lepidium sativum;
 Cucumis sativus; Phaseolus vulgaris; Synergism; Antagonism;
 Weed control
 
 Abstract:  Three allelochemicals from rye or its breakdown
 products were evaluated for activity on garden cress (Lepidum
 sativum L.), barnyardgrass [Echinochloa crus-galli (L.)
 Beauv.], cucumber (Cucumis sativus L.), and snap bean
 (Phaseolus vulgaris L.). 2,4-Dihydroxy-1,4(2H)-benzoxazin-3-
 one (DIBOA), 2(3H)-benzoxazolinone (BOA), and 2,2'-oxo-1,1'-
 azobenzene (AZOB) were all applied singly at 50, 100, and 200
 ppm and in two- and three-way combinations each at 50 and 100
 ppm. AZOB at 100 and 200 ppm produced 38-49% more inhibition
 than DIBOA, while combinations of BOA/DIBOA, which contained
 AZOB at 100 ppm had 54-90% more inhibition when compared to
 DIBOA/BOA combinations. All combinations were slightly
 antagonistic to barnyardgrass, while several combinations
 caused a synergistic response to garden cress germination and
 growth. Cucumbers and snap beans exhibited both types of
 responses, depending on the allelochemical combination and
 application rate. The plant-produced benzoxazinones were more
 inhibitory to crops than weeds. Therefore, improved herbicidal
 selectivity would be expected if there were rapid
 transformation of the benzoxazinones to the microbially
 produced AZOB.
 
 
 5                                  NAL Call. No.: QD415.A1J6
 6,10,14-Trimethylpentadecan-2-one: a Bermuda grass
 phagostimulant to fall armyworm (Lepidoptera: Noctuidae).
 Mohamed, M.A.; Quisenberry, S.S.; Moellenbeck, D.J.
 New York, N.Y. : Plenum Press; 1992 Apr.
 Journal of chemical ecology v. 18 (4): p. 673-682; 1992 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Spodoptera frugiperda; Cultivars; Cynodon
 dactylon; Allelopathy; Feeding behavior; Plant composition;
 Isoprenoids; Ketones; Bioassays; Phagostimulants; Insect
 control
 
 Abstract:  A phagostimulant, 6,10,14-trimethylpentadecan-2-one
 (phytone), was isolated and identified from Bermuda grass,
 Cynodon dactylon (L.). The phagostimulant activity of this
 isoprenoid ketone was established from bioassays of fall
 armyworm larvae, Spodoptera frugiperda (J.E. Smith). Larvae
 displayed increased body mass accumulation as well as
 preference to diet supplemented with this molecule. Neonate
 larvae fed diet supplemented with chromatographic isolates of
 phytone-containing fractions from six Bermuda grass cultivars
 showed a 10-40% increase in body mass accumulation as compared
 with controls. This variation in larval body mass accumulation
 seems attributable to a differential concentration of phytone
 in the cultivars, which ranged from 0.5 to 43 ppm.
 Additionally, first-instar larvae responded preferentially to
 diet pellets topically treated with phytone in concentrations
 as low as 0.1 ppm.
 
 
 6                                     NAL Call. No.: 450 R11
 The action of steroidal alkaloids on the ground meristem
 tissue of the root axis of lettuce seedlings.
 Ghazi, M.; Myers, G.A.
 Oxford : Pergamon Journals; 1990 Apr.
 Environmental and experimental botany v. 30 (2): p. 235-242.
 ill; 1990 Apr. Includes references.
 
 Language:  English
 
 Descriptors: Solanaceae; Weeds; Lactuca sativa; Phytotoxicity;
 Alkaloids; Solanine; Necroses (plant); Histopathology; Roots;
 Growth rate; Root meristems; Allelopathy
 
 
 7                                  NAL Call. No.: QD241.K453
 Alanto- and isoalantolactones.
 Milman, I.A.
 New York, N.Y. : Consultants Bureau; 1990 Nov.
 Chemistry of natural compounds v. 26 (3): p. 251-262; 1990
 Nov.  Translated from: Khimiia Prirodnykh Soedinenii, v. 26
 (3), 1990, p. 307-320. (QD241.K45).  Literature review. 
 Includes references.
 
 Language:  English; Russian
 
 Descriptors: U.S.S.R.; Sesquiterpenoid lactones; Plant
 composition; Isolation; Inula; Physicochemical properties;
 Allelopathy
 
 
 8                                 NAL Call. No.: QK938.F6C32
 Alelopatia v lesnych ekosystemoch  [Allelopathy in forest
 ecosystems]. Caboun, Vladimir
 Bratislava : VEDA, vydavatel'stvo Slovenskej akademie vied,
 1990; 1990. 118 p. : ill. ; 24 cm. (Biologicke prace). 
 Summaries in Russian and English. Includes bibliographical
 references (p. 99-109).
 
 Language:  Slovak
 
 Descriptors: Forest ecology; Allelopathy
 
 
 9                                    NAL Call. No.: 421 J822
 Allelochemical content of selected popcorn silks: effects on
 growth of corn earworm larvae (Lepidoptera: Noctuidae).
 Wiseman, B.R.; Snook, M.E.; Wilson, R.L.; Isenhour, D.J.
 Lanham, Md. : Entomological Society of America; 1992 Dec.
 Journal of economic entomology v. 85 (6): p. 2500-2504; 1992
 Dec.  Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Tassels; Allelochemicals; Antibodies;
 Pest resistance; Helicoverpa zea; Larvae; Growth; Weight
 
 Abstract:  Silks of popcorn, zea mays L., in the Eldredge
 collection were evaluated for effects on growth of larvae of
 the corn earworm, Helicoverpa zea (Boddie), and assayed for
 content of maysin, chlorogenic acid, apimaysin, and 3'-
 methoxymaysin. The weights of 9-d-old larvae ranged from 8 mg
 on "PI340856" to 952 mg on "PI340855". Silks from about one-
 third of the popcorn collection produced corn earworm larvae
 equal to or smaller than larvae that fed on silks of the
 resistant standard, "Zapalote Chico". Content of maysin, a
 luteolin-C-glycoside, ranged from zero for eighteen PIs to
 1.128% for "PI340856". Significant negative correlations were
 found between contents of all four allelochemicals assayed and
 larval weights. Silks of "PI340853" had no detectable levels
 of any of the allelochemicals assayed, yet they produced
 larvae with an average weight of only 37 mg. Silks of several
 introductions had higher concentrations of chlorogenic acid,
 apimaysin, and 3'-methoxymaysin than silks of "Zapalote
 Chico".
 
 
 10                                 NAL Call. No.: QD415.A1J6
 An allelochemical elicits arrestment in Apanteles kariyai in
 feces of nonhost larvae Acantholeucania loreyi.
 Takabayashi, J.; Takahashi, S.
 New York, N.Y. : Plenum Press; 1990 Jun.
 Journal of chemical ecology v. 16 (6): p. 2009-2017; 1990 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Apanteles; Noctuidae; Mythimna separata;
 Allelochemicals; Oviposition; Interactions; Insect control;
 Biological control
 
 Abstract:  Females of the larval parasitoid Apanteles kariyai
 (Watanabe) (Hymenoptera: Braconidae) are arrested on fecal
 pellets of Acantholeucania loreyi (Duponchel) (Lepidoptera:
 Noctuidae) larvae. Upon subsequent antennal contact with an A.
 loreyi larva, females sting it with their ovipositor. However,
 such stinging did not result in any offspring. The
 allelochemical involved in feces has been identified and is
 identical to a kairomone of the host Pseudaletia separata
 (Lepidoptera: Noctuidae). In contrast to A. loreyi, P.
 separata is a suitable host for A. kariyai, and oviposition in
 P. separata results in offspring production. The
 allelochemical mediating the interaction between A. loreyi and
 A. kariyai is discussed in the context of current
 allelochemical terminology.
 
 
 11                                 NAL Call. No.: QD415.A1J6
 Allelochemical regulation of reproduction and seed germination
 of two Brazilian Baccharis species by phytotoxic
 trichothecenes. Kuti, J.O.; Jarvis, B.B.; Mokhtari-Rejali, N.;
 Bean, G.A. New York, N.Y. : Plenum Press; 1990 Dec.
 Journal of chemical ecology v. 16 (12): p. 3441-3453; 1990
 Dec.  Includes references.
 
 Language:  English
 
 Descriptors: Baccharis cordifolia; Baccharis megapotamica;
 Allelopathy; Plant composition; Phytotoxins; Seed germination;
 Trichothecenes; Pollination; Allelochemicals
 
 Abstract:  The potent phytotoxic trichothecene roridins and
 baccharinoids occur naturally in the Brazilian plants,
 Baccharis coridifolia and B. megapotamica. Biosynthesis of
 roridins in B. coridifolia appears to be linked to
 pollination, and the phytotoxins then accumulate in the seed.
 The roles of the phytotoxins in pollination, seed maturation,
 and germination of the Baccharis species were investigated.
 The high production of roridins occurred only in seeds
 resulting from intraspecific pollination, and the
 concentration of the toxins in the seeds generally increased
 with seed maturity. Removal of seed coats from trichothecene-
 producing Brazilian Baccharis species (B. coridifolia and B.
 megapotamica) and non-trichothecene-producing American
 Baccharis species (B. halimifolia and B. glutinosa) resulted
 in improved seed germination of B. halimifolia and B.
 glutinosa but complete inhibition of seed germination of B.
 coridifolia and B. megapotamica. Addition of seed coat
 extracts of the Brazilian Baccharis species of dilute
 solutions (10(-6) micrograms/ml) of roridins or baccharinoids
 to the decoated seeds of B. coridifolia and B. megapotamica
 resulted in germination, while seeds of B. halimifolia and B.
 glutinosa were killed by the phytotoxins. Roridins interacted
 with gibberellic acid, a germination promoter, but not with
 abscisic acid, a germination inhibitor. The results from this
 study suggest that macrocyclic tricothecenes have a regulatory
 role(s) on reproduction and germination of Brazilian Baccharis
 species in their natural habitat.
 
 
 12                                 NAL Call. No.: QD415.A1J6
 Allelochemicals from Polygonum sachalinense Fr. Schm.
 (Polygonaceae). Inoue, M.; Nishimura, H.; Li, H.H.; Mizutani,
 J.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Oct.
 Journal of chemical ecology v. 18 (10): p. 1833-1840; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Root exudates; Plant composition; Anthraquinones;
 Allelopathy; Allelochemicals; Bioassays; Growth; Inhibition;
 Seedlings; Weed control
 
 Abstract:  The root exudates from Polygonum sachalinense in a
 recirculating system significantly inhibited lettuce seedling
 growth. The rhizomes and roots of P. sachalinense were
 extracted with 80% acetone. Bioassay of the neutral-acidic
 fraction on the TLC agar plate showed the inhibitory activity
 corresponded to the two yellow pigment bands. Two orange
 needles were isolated and identified as anthraquinone
 compounds: emodin and physcion. Both compounds exhibited
 inhibitory activities against the seedling growth of several
 testing plant species. Glucosides were isolated from P.
 sachalinense and were identified as emodin-1-O-beta-D-
 glucoside and physcion-1-O-beta-D-glucoside, respectively. On
 plant growth bioassay, these glucosides showed no phytotoxic
 activity against lettuce seedlings. The concentrations of
 emodin, physcion, and their glucosides from rhizome with
 roots, aerial parts, fallen leaves, and soil were determined.
 The rhizome with roots and fallen leaves contained emodin and
 physcion at relatively high concentrations. Emodin also occurs
 in the soil of this plant community with effective
 concentrations in the fall. The results indicate that these
 anthraquinones are responsible for the observed interference
 and are potent allelopathic substances.
 
 
 13                                 NAL Call. No.: QD415.A1J6
 Allelochemicals in foliage of unfavored tree hosts of the
 gypsy moth, Lymantria dispar L. 1. Alkaloids and other
 components of Liriodendron tulipifera L. (Magnoliaceae), Acer
 rubrum L. (Aceraceae), and Cornus florida L. (Cornaceae).
 Barbosa, P.; Gross, P.; Provan, G.J.; Pacheco, D.Y.; Stermitz,
 F.R. New York, N.Y. : Plenum Press; 1990 May.
 Journal of chemical ecology v. 16 (5): p. 1719-1730; 1990 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lymantria dispar; Antifeedants; Liriodendron
 tulipifera; Acer rubrum; Cornus florida; Leaves; Plant
 composition; Alkaloids; Sesquiterpenoid lactones
 
 Abstract:  Early theories on plant chemical defense against
 herbivory emphasized that apparent and unapparent plants were
 primarily defended by different types of compounds. More and
 more evidence suggests that both quantitative and qualitative
 defenses are found in apparent plants and that they can play a
 defensive role against herbivores. A survey of the literature
 on the gypsy moth suggests not only that there is a large
 variety of qualitative compounds, as well as the expected
 quantitative ones, but that unfavored hosts of the gypsy moth
 are associated with the presence of alkaloids. Foliage of
 three tree species, Liriodendron tulipifera L., Acer rubrum
 L., and Cornus florida L., was examined to confirm the
 presence of alkaloids and other major secondary metabolites.
 The known sesquiterpene lactone, lipiferolide, and the sugar
 derivative, liriodendritol, were components of L. tulipifera
 leaves, along with a bisphenylpropanoid previously found only
 in nutmeg. Alkaloid content [i.e., (-)-N-methylcrotsparine
 content] was low and leaves tested positive for HCN. Leaves of
 A. rubrum L. were examined for the presence of gramine, but
 none could be detected. No alkaloids were detected in Cornus
 florida.
 
 
 14                                 NAL Call. No.: QD415.A1J6
 Allelochemicals in foliage of unfavored tree hosts of the
 gypsy moth, Lymantria dispar L. 2. Seasonal variation of
 saponins in Ilex opaca and identification of saponin
 aglycones.
 Barbosa, P.; Gross, P.; Provan, G.J.; Stermitz, F.R.
 New York, N.Y. : Plenum Press; 1990 May.
 Journal of chemical ecology v. 16 (5): p. 1731-1738; 1990 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lymantria dispar; Antifeedants; Ilex opaca;
 Leaves; Plant composition; Saponins; Pest resistance; Insect
 control; Biological control
 
 Abstract:  A greater variety of qualitative chemical defenses
 has been reported in eastern forest trees than might be
 expected from current interpretation of the plant apparency
 theory. For the gypsy moth there is an association between the
 occurrence of alkaloids and unfavorability of certain tree
 species, as well as the presence of saponins. The latter
 association, however, is not statistically significant.
 Species in the genus Ilex have been reported to contain both
 alkaloids and saponins (Barbosa and Krischick, 1987). In this
 study, determinations were made of the occurrence of alkaloids
 and saponins in I. opaca and their changes in concentration
 over time. No alkaloids were detected. Saponins were isolated,
 and the aglycone siaresinolic acid was identified. Saponin
 concentration changes seasonally, being highest in early May
 and lowest in early June leaves.
 
 
 15                                  NAL Call. No.: SF601.A47
 Allelochemicals in plant foods and feedingstuffs. 1.
 Nutritional, biochemical and physiopathological aspects in
 animal production.
 Aletor, V.A.
 Manhattan, Kan. : Kansas State University; 1993 Feb.
 Veterinary and human toxicology v. 35 (1): p. 57-67; 1993 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Feeds; Antinutritional factors; Allomones;
 Literature reviews
 
 
 16                                 NAL Call. No.: QD415.A1J6
 Allelochemicals in soil from no-tillage versus conventional-
 tillage wheat (Triticum aestivum) fields.
 Cast, K.G.; McPherson, J.K.; Pollard, A.J.; Krenzer, E.G. Jr;
 Waller, G.R. New York, N.Y. : Plenum Press; 1990 Jul.
 Journal of chemical ecology v. 16 (7): p. 2277-2289; 1990 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Triticum aestivum; No-tillage; Tillage;
 Allelochemicals; Wheat soils; Chemical composition; Bioassays;
 Fatty acids; Roots; Growth
 
 Abstract:  Putative allelochemicals found in the soil of no-
 tillage and conventional-tillage wheat plots near Stillwater,
 Oklahoma, were obtained by a mild alkaline aqueous extraction
 procedure, bioassayed to determine their biological activity,
 purified, and analyzed with a capillary gas chromatography-
 mass spectrometry-data analysis system. The most significant
 inhibition was found in bioassays of extracts from soil
 collected immediately after harvest in June, July, and August.
 No-tillage soils produced significant inhibition during the
 rest of the year also. Mass spectrometry showed fatty acids as
 the most abundant compounds. However, when bioassayed
 authentic samples of the five free fatty acids showed no
 significant biological activity toward wheat.
 
 
 17                                 NAL Call. No.: QD415.A1J6
 Allelochemicals produced during glucosinolate degradation in
 soil. Brown, P.D.; Morra, M.J.; McCaffrey, J.P.; Auld, D.L.;
 Williams, L. III New York, N.Y. : Plenum Press; 1991 Oct.
 Journal of chemical ecology v. 17 (10): p. 2021-2034; 1991
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Brassica napus; Oilseeds; Plant composition;
 Allelochemicals; Glucosinolates; Degradation; Thiocyanates;
 Quantitative analysis; Insect control; Biological control
 
 Abstract:  A variety of plant pests are suppressed by the
 incorporation of cruciferous plant material into soil.
 Although this effect is attributed to decomposition of
 glucosinolates into toxic products, little is known concerning
 glucosinolate degradation in the soil environment. Arenas (30
 X 18 X 8 cm) that contained soil amended with 30 g defatted
 winter rapeseed meal (Brassica napus L.)/kg soil on one half
 and unamended soil on the other were constructed.
 Isothiocyanate concentrations in the soil were measured using
 infrared analysis of CCl4 extracts, and ionic thiocyanate
 (SCN-) using ion chromatography on aqueous extracts.
 Quantities were monitored during a 100-hr time period in
 conjunction with a wireworm bioassay. Isothiocyanate
 production reached a maximum of 301 nmol/g soil at 2 hr, but
 decreased by 90% within 24 hr. Production of SCN- reached a
 maximum of 180 nmol/g soil at 8 hr but persisted longer than
 isothiocyanate. Separate late instar wireworms (Limonius
 infuscatus Mots.) were repelled by the presence of rapeseed
 meal in less than 24 hr even though the meal was shown in
 separate experiments not to be toxic. We propose that rapidly
 produced isothiocyanates are responsible for this repellency,
 but other products such as SCN- may play a role.
 
 
 18                                 NAL Call. No.: QD415.A1J6
 Allelopathic activity in wheat-conventional and wheat-no-till
 soils: development of soil extract bioassays.
 Blum, U.; Gerig, T.M.; Worsham, A.D.; Holappa, L.D.; King,
 L.D. New York, N.Y. : Plenum Publishing Corporation; 1992 Dec.
 Journal of chemical ecology v. 18 (12): p. 2191-2221; 1992
 Dec.  Includes references.
 
 Language:  English
 
 Descriptors: North Carolina; Cabt; Triticum aestivum; Glycine
 max; Allelopathy; Germination; Bioassays; Phenolic acids;
 Hydroxamic acids; Soil analysis; Extraction; Tillage;
 Pharbitis hederacea; Trifolium incarnatum; Weed control
 
 Abstract:  The primary objective of this research was to
 determine if soil extracts could be used directly in bioassays
 for the detection of allelopathic activity. Here we describe:
 (1) a way to estimate levels of allelopathic compounds in
 soil; (2) how pH, solute potential, and/or ion content of
 extracts may modify the action of allelopathic compounds on
 germination and radicle and hypocotyl length of crimson clover
 (Trifolium incarnatum L.) and ivy-leaved morning glory
 (Ipomoea hederacea L. Jacquin.), and 3) how biological
 activity of soil extracts may be determined. A water-autoclave
 extraction procedure was chosen over the immediate-water and
 5-hr EDTA extraction procedures, because the autoclave
 procedure was effective in extracting solution and reversibly
 bound ferulic acid as well as phenolic acids from wheat
 debris. The resulting soil extracts were used directly in
 germination bioassays. A mixture of phenolic acids similar to
 that obtained from wheat-no-till soils did not affect
 germination of clover or morning glory and radicle and
 hypocotyl length of morning glory. The mixture did, however,
 reduce radicle and hypocotyl length of clover. Individual
 phenolic acids also did not inhibit germination, but did
 reduce radicle and hypocotyl length of both species. 6-MBOA
 (6-methoxy-2,3-benzoxazolinone), a conversion product of 2-o-
 glucosyl-7-methoxy-1,4-benzoxacin-3-one, a hydroxamic acid in
 living wheat plants, inhibited germination and radicle and
 hypocotyl length of clover and morning glory. 6-MBOA, however,
 was not detected in wheat debris, stubble, or soil extracts.
 Total phenolic acids (FC) in extracts were determined with
 Folin and Ciocalteu's phenol reagent. Levels of FC in wheat-
 conventional-till soil extracts were not related to
 germination or radicle and hypocotyl length either species.
 Levels of FC in wheat-no-till soil extracts were also not
 related to germination of clover or morning glory, but were
 inversely related to radicle and hypocotyl length of clover
 and morning glory. FC values, solute potential, and acidity of
 wheat-no-till soil extracts appeared to be independent
 (additive) in action on clover radicle and hypocotyl length.
 Radicle and hypocotyl length of clover was inversely related
 to increasing FC and solute potential and directly related to
 decreasing acidity. Biological activity of extracts was
 determined best from slopes of radicle and hypocotyl length
 obtained from bioassays of extract dilutions, Thus, data
 derived from the water-autoclave extraction procedure, FC
 analysis, and slope analysis for extract activity in
 conjunction with data on extract pH and solute potential can
 be used to estimate allelopathic activity of wheat-no-till
 soils
 
 
 19                                  NAL Call. No.: RS160.I47
 Allelopathic activity of the essential oils of Nigerian
 medicinal plants. Oguntimein, B.O.; Elakovich, S.D.
 Lisse, Netherlands : Swets & Zeitlinger; 1991 Feb.
 International journal of pharmacognosy v. 29 (1): p. 39-44;
 1991 Feb. Includes references.
 
 Language:  English
 
 Descriptors: Nigeria; Eugenia uniflora; Piper guineense;
 Chromolaena; Medicinal plants; Lactuca sativa; Plant extracts;
 Essential oils; Allelopathins; Allelopathy; Seedling growth;
 Bioassays
 
 
 20                                 NAL Call. No.: QD415.A1J6
 Allelopathic and autotoxic effects of Anastatica hierochuntica
 L. Hegazy, A.K.; Mansour, K.S.; Abdel-Hady, N.F.
 New York, N.Y. : Plenum Press; 1990 Jul.
 Journal of chemical ecology v. 16 (7): p. 2183-2193; 1990 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cruciferae; Desert plants; Allelopathy; Plant
 composition; Plant extracts; Bioassays; Seedling growth; Seed
 germination; Cell division; Inhibition
 
 Abstract:  Laboratory experiments were undertaken to
 investigate the autotoxic effects of Anastatica hiertochuntica
 and possible effects on five other desert plants: Rumex
 cyprius, Trigonella stellata, Diplotaxis harra, Cleome
 droserifolia, and Farsetia aegyptia. Seed germination.
 seedling growth, and cell division of all species tested were
 inhibited by the shoot aqueous extract of A. hierochuntica. A
 gradual increase in the percentage of prophase and decrease in
 the other mitotic stages as well as the mitotic index were
 observed with increasing extract concentration. At an extract
 concentration of 8% the mitotic index was reduced from the
 control by 55% in C. droserifolia, 54% in T. stellata, 45% in
 F. aegyptia, 43% in A. hierochuntica, and 35% in R. cyprius.
 The inhibitory substances are apparently released onto soil by
 repeated washing of the standing plants by rain and dew
 interception.
 
 
 21                                   NAL Call. No.: 450 AM36
 Allelopathic and herbicidal effects of extracts from tree of
 heaven (Ailanthus altissima).
 Heisey, R.M.
 Columbus, Ohio : Botanical Society of America; 1990 May.
 American journal of botany v. 77 (5): p. 662-670. ill; 1990
 May.  Includes references.
 
 Language:  English
 
 Descriptors: New York; Ailanthus altissima; Lepidium sativum;
 Radicles; Growth rate; Phytotoxicity; Plant extracts;
 Herbicidal properties; Allelopathy; Allelopathins; Seasonal
 variation; Wood; Bark; Seeds; Leaves
 
 
 22                               NAL Call. No.: 79.9 SO8 (P)
 Allelopathic cover crops to reduce herbicide input.
 Worsham, A.D.
 Raleigh, N.C. : The Society :.; 1991.
 Proceedings - Southern Weed Science Society v. 44: p. 58-69;
 1991.  Paper presented at the meeting on "Perception: Fact or
 Fiction", held January 14-16, 1991, San Antonio, Texas. 
 Includes references.
 
 Language:  English
 
 Descriptors: North Carolina; Cover crops; Allelopathy;
 Herbicides; Application rates; Weed control
 
 
 23                                 NAL Call. No.: QD415.A1J6
 Allelopathic dominance of Miscanthus transmorrisonensis in an
 alpine grassland community in Taiwan.
 Chou, C.H.; Lee, Y.F.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2267-2281; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Taiwan; Miscanthus transmorrisonensis;
 Allelopathy; Grasslands; Plant communities; Phenolic
 compounds; Phytotoxins
 
 Abstract:  A study site located at 2600 m elevation in
 Tartarchia Anpu, Nantou county, Taiwan, exhibits a unique
 grassland community composed of two principal species.
 Miscanthus transmorrisonensis and Yushinia niitakayamensis,
 and 35 other species. The relative frequencies of the two
 species are 12% and 11%, while their relative coverages ate
 25% and 19.5%, respectively. The values for the remaining 35
 species are lower than 4% each, while species diversity of the
 community is -3.04839, indicating great diversity. To
 elucidate the mechanism of dominance of M. transmorrisonensis,
 allelopathic evaluation of the plant was conducted. Aqueous
 extracts of M. transmorrisonensis plant parts with two
 ecotypes were bioassayed. The extracts showed significant
 phytotoxic effects on seed germination and radicle growth of
 four tested plants: rye grass, lettuce, and two varieties of
 Chinese cabbage. In addition, rhizosphere soils under
 Miscanthus also exhibited significant phytotoxicity,
 indicating that allelopathic interaction was involved. Some
 responsible phytotoxic phenolics, namely, p-coumaric, ferulic,
 vanillic, protocatechuic, o-hydroxyphenylacetic, and m-
 hydroxyphenylacetic acids, and 4-hydroxycoumarin and
 phloridzin were identified. Allelopathy thus can play an
 important role in regulating plant diversity in the field.
 
 
 24                                    NAL Call. No.: QK1.A28
 Allelopathic effect of Harpullia imbricata Thw. leaf leachate
 on Phaseolus mungo L.
 Xavier, A.
 Meerut, India : Society for Advancement of Botany; 1990 Dec.
 Acta botanica Indica v. 18 (2): p. 293-295; 1990 Dec. 
 Includes references.
 
 Language:  English
 
 Descriptors: Vigna mungo; Sapindaceae; Allelopathins; Plant
 extracts; Leaves; Leachates; Seed germination; Germination
 inhibitors; Growth rate; Inhibition; Protein synthesis; Amino
 acids; Proteolysis
 
 
 25                                  NAL Call. No.: 450 P5622
 Allelopathic effect of hydroxamic acids from cereals on Avena
 sativa and A. fatua.
 Perez, F.J.
 Oxford : Pergamon Press; 1990.
 Phytochemistry v. 29 (3): p. 773-776; 1990.  Includes
 references.
 
 Language:  English
 
 Descriptors: Triticum durum; Avena sativa; Avena fatua; Crop
 plants as weeds; Weed control; Allelopathy; Biological
 control; Hydroxamic acids; Seed germination; Growth rate
 
 
 26                                NAL Call. No.: S596.53.S69
 Allelopathic effect of sweetpotato (Ipomoea batatas) cultivars
 on certain weed and vegetable species.
 Reinhardt, C.F.; Meissner, R.; Nel, P.C.
 Pretoria : Bureau for Scientific Publications, Foundation for
 Education, Science and Technology, [1984-; 1993 Feb.
 South African journal of plant and soil; Suid-Afrikaanse
 tydskrif vir plant en grond v. 10 (1): p. 41-44; 1993 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Ipomoea batatas; Cultivars; Allelopathy
 
 
 27                                   NAL Call. No.: 470 C16C
 Allelopathic effects by Empetrum hermaphroditum on development
 and nitrogen uptake by roots and mycorrhizae of Pinus
 silvestris.
 Nilsson, M.C.; Hogberg, P.; Zackrisson, O.; Fengyou, W.
 Ottawa, Ont. : National Research Council of Canada; 1993 Apr.
 Canadian journal of botany; Journal canadien de botanique v.
 71 (4): p. 620-628; 1993 Apr.  Includes references.
 
 Language:  English
 
 Descriptors: Empetrum; Pinus sylvestris; Paxillus involutus;
 Allelopathy; Plant extracts; Nitrogen; Nutrient uptake; Roots;
 Ectomycorrhizas; Growth; Dry matter accumulation; Root tips;
 Seedling growth; Root shoot ratio
 
 
 28                                   NAL Call. No.: SD13.C35
 Allelopathic effects by Empetrum hermaphroditum on seed
 germination of two boreal tree species.
 Zackrisson, O.; Nilsson, M.C.
 Ottawa, Ont. : National Research Council of Canada; 1992 Sep.
 Canadian journal of forest research; Revue canadienne de
 recherche forestiere v. 22 (9): p. 1310-1319; 1992 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Sweden; Pinus sylvestris; Populus tremula;
 Allelopathy; Allelopathins; Seedgermination; Empetrum; Forest
 litter; Plant secretions; Soil flora; Metabolic
 detoxification; Humus; Boreal forests
 
 Abstract:  Indoor and outdoor experiments demonstrated that
 allelopathy is an important factor explaining seed
 regeneration failures of Scots pine (Pinus silvestris L.) in
 forest floor vegetation dominated by the dwarf shrub Empetrum
 hermaphroditum Hagerup. Scanning electron micrograph views of
 the leaf surfaces of E. hermaphroditum reveal secretory glands
 that are shown to be involved in the release of water-soluble
 phytotoxic substances. Bioassays indicate that low doses and
 short exposure times of seeds to leachates have strong
 negative effects on germination and early root development.
 Activated carbon can eliminate the inhibitory effects of
 leachates and litter. This technique demonstrates the
 occurrence of allelopathic interference by E. hermaphroditum
 on seed germination of both Scots pine and aspen (Populus
 tremula L.). In a field experiment the allelopathic effects by
 E. hermaphroditum are strong during early spring when
 germination and growth initiate and ground ice still is
 present. Extracts passed through soils collected from an E.
 hermaphroditum site were detoxified. while those passed
 through sterilized soil were not. Therefore, microorganisms
 may detoxify the allelochemicals under some environmental
 conditions.
 
 
 29                                     NAL Call. No.: SB1.H6
 Allelopathic effects of alfalfa plant residues on emergence
 and growth of cucumber seedlings.
 Ells, J.E.; McSay, A.E.
 Alexandria, Va. : American Society for Horticultural Science;
 1991 Apr. HortScience v. 26 (4): p. 368-370; 1991 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Medicago sativa; Plant residues; Allelopathy;
 Phytotoxicity; Cucumis sativus; Seed germination; Seedling
 growth; Growing media
 
 Abstract:  Growth chamber tests demonstrated that alfalfa
 (Medicago sativa L.) residue is toxic to cucumber (Cucumis
 sativus L.) seed germination and seedling growth. Ground
 alfalfa roots at 0.5% (w/w, dry weight) inhibited germination
 when added to the growing medium. Alfalfa roots at 0.5% were
 also toxic to pregerminated cucumber seed. However, cucumber
 seedlings grew normally if this same medium was watered and
 incubated for > 1 day before planting. Alfalfa particle size
 in media influenced cucumber performance, with the
 intermediate size (1 to 2 mm) being lethal to cucumbers.
 
 
 30                                    NAL Call. No.: SD1.I54
 Allelopathic effects of Eucalyptus tereticornis on Phaseolus
 vulgaris seedlings.
 Puri, S.; Khara, A.
 Oxon : A B Academic; 1991.
 The International tree crops journal v. 6 (4): p. 287-293;
 1991.  Includes references.
 
 Language:  English
 
 Descriptors: Phaseolus vulgaris; Eucalyptus tereticornis;
 Seedlings; Allelopathy; Leaves; Bark; Seed germination; Roots;
 Shoots; Plant development; Leachates
 
 
 31                                    NAL Call. No.: SD1.I54
 Allelopathic effects of Parthenium hysterophorus on
 germination and seedling growth of a few multi-purpose trees
 and arable crops.
 Swaminathan, C.; Vinaya Rai, R.S.; Suresh, K.K.
 Oxon : A B Academic; 1990.
 The International tree crops journal v. 6 (2/3): p. 143-150;
 1990.  Includes references.
 
 Language:  English
 
 Descriptors: Acacia leucophloea; Casuarina equisetifolia;
 Eucalyptus tereticornis; Leucaena leucocephala; Allelopathy;
 Parthenium hysterophorus; Seed germination; Seedlings; Growth;
 Lactones
 
 
 32                                NAL Call. No.: S592.7.A1S6
 Allelopathic effects of plant seeds on nitrification: effects
 on ammonium oxidizers.
 Kholdebarin, B.; Oertli, J.J.
 Exeter : Pergamon Press; 1992 Jan.
 Soil biology and biochemistry v. 24 (1): p. 59-64; 1992 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Quercus rubra; Quercus petraea; Quercus robur;
 Camellia sinensis; Seeds; Kernels; Testas; Cotyledons;
 Powders; Leaves; Plant extracts; Soil bacteria; Nitrification
 inhibitors; Phenolic compounds; Nitrification; Ammonium;
 Oxidation; Biological activity in soil; Nitrites; Nitrate;
 Ammonium nitrogen; Nitrate nitrogen; Immobilization; Nitrogen
 fixation; Chemical reactions; Allelopathy
 
 Abstract:  Effects of cotyledon powder from seeds of higher
 plants (tea and several varieties of oaks) known to be rich in
 phenolic compounds on biological oxidation of NH(+4) and
 NO(-2) to NO(-3) were investigated. Treating culture solutions
 with cotyledon powder resulted in a rapid disappearance of
 both NH(+4)-N and NO(-2)-N during the first 2-3 days of the
 experiments. Such losses were believed to be due to fixation
 of NH(+4) and volatilization of NO(-2)-N by phenolic
 substances and also to reactions with other organic C
 compounds present in cotyledon tissues. It is also suggested
 that some of the NH(+4)-N and 4 NO(-2)-N may have been
 immobilized by heterotrophic bacterial growth.
 
 
 33                                NAL Call. No.: S592.7.A1S6
 Allelopathic effects of plant seeds on nitrification: effects
 on nitrite oxidizers.
 Kholdebarin, B.
 Exeter : Pergamon Press; 1992 Jan.
 Soil biology and biochemistry v. 24 (1): p. 65-69; 1992 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Camellia sinensis; Quercus robur; Seeds;
 Cotyledons; Powders; Soil bacteria; Glucose; Cell cultures;
 Nutrient solutions; Sodium nitrite; Oxidation; Nitrate;
 Nitrification; Fixation; Organic compounds; Volatilization;
 Nitrogen; Ammonium; Allelopathy; Immobilization
 
 Abstract:  Effects of cotyledon powder, from tea and oak
 seeds, on oxidation of NO(-2) to NO(-3) in nitrification were
 investigated. Presence of cotyledon powder or glucose in
 culture solutions greatly stimulated the rapid disappearance
 of NO(-2) from solutions. However, the amount of NO(-3) as the
 end product of nitrification was drastically reduced in the
 presence of seed cotyledon or glucose. Based on the results
 obtained from double enrichment experiments and also from
 experiments done with sterile and non-sterile soil-free
 solutions, it was concluded that the decrease in the amount of
 NO(-3) in nitrification seems to be due to fixation,
 volatilization and immobilization of nitrogen by organic
 substances present in ground cotyledons of tea and oak seeds;
 direct effects on nitrifying organisms seem to be negligible.
 
 
 34                                     NAL Call. No.: S1.T49
 Allelopathic effects of two grasses on seed germination of
 three wildlife food plants.
 Fulbright, N.; Fulbright, T.E.
 Canyon, Tex. : The Consortium; 1990.
 Texas journal of agriculture and natural resources : a
 publication of the Agricultural Consortium of Texas v. 4: p.
 31-32; 1990.  Includes references.
 
 Language:  English
 
 Descriptors: Texas; Colinus Virginianus; Sorghum almum;
 Panicum coloratum; Panicum antidotale; Cenchrus ciliaris;
 Dichanthium annulatum; Allelopathy; Seed germination;
 Leachates; Wildlife management
 
 
 35                                 NAL Call. No.: QD415.A1J6
 Allelopathic effects of water extracts of Artemisia princeps
 var. orientalis on selected plant species.
 Kil, B.S.; Yun, K.W.
 New York, N.Y. : Plenum Press; 1992 Jan.
 Journal of chemical ecology v. 18 (1): p. 39-51; 1992 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Artemisia princeps; Allelopathy; Leaves; Stems;
 Roots; Extracts; Bioassays; Seed germination; Seedlings
 
 Abstract:  The allelopathic effects of wormwood plants
 (Artemisia princeps var. orientalis) and their possible
 phytotoxicity on receptor species were investigated. The
 aqueous extracts of mature leaf, stem, and root of wormwood
 plants caused significant inhibition in germination and
 decreased seedling elongation of receptor plants, whereas
 germination of some species was not inhibited by extracts of
 stems and roots. Dry weight growth was slightly increased at
 lower concentrations of the extract, whereas it was
 proportionally inhibited at higher concentrations. The calorie
 value of the organic matter in receptor plants measured by
 bomb calorimeter was reduced proportionally to the extract
 concentration. However, results with extracts of juvenile leaf
 did not correlate with inhibition or promotion of elongation
 and dry weight.
 
 
 36                                 NAL Call. No.: QD415.A1J6
 Allelopathic inhibition of Cynodon dactylon (L.) Pers. and
 other plant species by Euphorbia prostrata L.
 Alsaadawi, I.S.; Sakeri, F.A.K.; Al-Dulaimy, S.M.
 New York, N.Y. : Plenum Press; 1990 Sep.
 Journal of chemical ecology v. 16 (9): p. 2747-2754; 1990 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Euphorbia prostrata; Allelopathy; Cynodon
 dactylon; Soil analysis; Bioassays; Biological control
 
 Abstract:  Field observations indicated that Euphorbia
 prostrata strongly interferes with Cynodon dactylon (L.) Pers.
 Analysis of some physical and chemical soil factors indicated
 that competition was not the dominant factor of that
 interference. Soil collected from under E. prostrata stands
 was very inhibitory to seed germination and seeding growth of
 some of the test species including C. dactylon. This suggests
 the presence of inhibitory compounds in soil of E. prostrata
 stands. Subsequent experiments showed that aqueous extract,
 decaying residues, and root exudates of E. prostrata were
 inhibitory to most of the test species including C. dactylon.
 Thus, it appears that allelopathy is the major component of
 the interference, with competition probably accentuating its
 effect. It also was found that allelopathy is an important
 component of the interference by E. prostrata against
 Amaranthus retro-flexus, Medicago sativa, and Gossypium
 hirsutum.
 
 
 37                                  NAL Call. No.: 450 P5622
 Allelopathic inhibition of seed germination by Cinchona
 alkaloids?. Aerts, R.J.; Snoeijer, W.; Meijden, E. van der;
 Verpoorte, R. Oxford : Pergamon Press; 1991.
 Phytochemistry v. 30 (9): p. 2947-2951; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Cinchona; Catharanthus roseus; Rubiaceae; Ocimum
 Americanum; Seed germination; Germination inhibitors; Plant
 extracts; Alkaloids; Allelopathins; Roots; Growth inhibitors
 
 Abstract:  The inhibition of seed germination by quinoline
 alkaloids synthesized by plants of the tropical genus Cinchona
 was studied. The germination of Ocimum (a tropical herb), of
 Spermacoce and Catharanthus (two tropical, alkaloid-producing
 plants), and of Cinchona itself was strongly inhibited by the
 alkaloids when applied at concentrations higher than about 0.3
 mM. To test for the possible allelopathic significance of this
 finding, the soil in which two-year-old Cinchona plants were
 grown was examined for its quinoline alkaloid content.
 Although the roots of the plants contain high concentrations
 of these alkaloids (ca 10 mM), in the soil only very low
 concentrations were found (ca 0.02 mM). Upon germination of
 seeds sown close by the plants, no toxic effects were
 observed. So, although several studies have reported
 inhibition of seed germination by Cinchona alkaloids under
 laboratory conditions, our results indicate that this property
 does not play a role under natural circumstances at realistic
 concentrations.
 
 
 38                                 NAL Call. No.: QD415.A1J6
 Allelopathic potential of compounds isolated from Ipomoea
 tricolor Cav. (Convolvulaceae).
 Anaya, A.L.; Calera, M.R.; Mata, R.; Pereda-Miranda, R.
 New York, N.Y. : Plenum Press; 1990 Jul.
 Journal of chemical ecology v. 16 (7): p. 2145-2152; 1990 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Ipomoea tRicolor; Allelopathy; Plant composition;
 Plant extracts; Glycosides; Bioassays; Seeds; Seedling growth;
 Weed control; Amaranthus leucocarpus; Echinochloa crus-galli
 
 Abstract:  The allelopathic potential of I. tricolor, used in
 traditional agriculture as a weed controller, has been
 demonstrated by measuring the inhibitory activity of aqueous
 lixiviates and organic extracts of the plant material on
 seedling growth of Amaranthus leucocarpus and Echinochloa
 crusgalli. Bioactivity-guided fractionation of the most active
 organic extract led to the isolation of the allelopathic
 principles, which turned out to be a mixture of glycosides,
 having jalapinolic acid as the aglycone portion glycosidically
 linked in the 11 position to an oligosaccharide composed of
 glucose, rhamnose, and fucose, which also combines with the
 carboxyl group of the aglycone to form a macrocyclic ester.
 
 
 39                                 NAL Call. No.: QD415.A1J6
 Allelopathic potential of Nuphar lutea (L.) Sibth. & SM.
 (Nymphaeaceae). Elakovich, S.D.; Wooten, J.W.
 New York, N.Y. : Plenum Press; 1991 Apr.
 Journal of chemical ecology v. 17 (4): p. 707-714; 1991 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Nuphar lutea; Allelopathy; Bioassays; Lemna
 minor; Lactuca sativa; Osmotic pressure
 
 Abstract:  Aqueous extracts of Nuphar lurea (L.) Sibth. & Sm.
 leaves (blades plus petioles) and roots plus rhizomes were
 tested for allelopathic activity using lettuce seedling and
 Lemna minor L. assay systems. The 12.5. 25, 125, and 250 parts
 per thousand (ppt) treatments of both extracts killed the
 lettuce seedlings. At 2.5 ppt of extract, radicle growth of
 lettuce was 29% of the control for leaves and 31% of the
 control for roots plus rhizomes. Lemna minor frond number was
 reduced to 34% of the control by the 25 ppt leaf extract and
 to 43% of the control by the 25 ppt roots plus rhizomes
 extract. L. minor was killed by concentrations of 125 ppt and
 above of both plant part extracts. As expected, the frond
 number and total chlorophyll content measured by the L. minor
 assay were highly correlated. Osmotic potentials below 143
 MOsmol/kg had no influence on L. minor growth. Neither the
 osmotic potential nor the pH of the undiluted extracts of N.
 lutea were in the range known to influence the growth of
 either lettuce seedlings or L. minor. Nuphar lutea extracts
 were many times more inhibitory than 16 other hydrophytes we
 previously examined.
 
 
 40                                    NAL Call. No.: 18 J825
 Allelopathic potential of shoot and root leachates of certain
 weed species. Rani, M.S.; Babu, R.C.; Sheriff, M.M.; Perumal,
 R.K.P.
 Berlin, W. Ger. : Paul Parey; 1990.
 Zeitschrift fur Acker- und Pflanzenbau v. 164 (2): p. 81-84;
 1990.  Includes references.
 
 Language:  English
 
 Descriptors: Tamil nadu; Weeds; Shoots; Roots; Toxic exudates;
 Phenolic content; Phytotoxicity; Allelopathy; Sorghum bicolor;
 Vigna mungo
 
 
 41                                   NAL Call. No.: 470 C16C
 Allelopathic potential of western coneflower (Rudbeckia
 occidentalis). Ferguson, D.E.
 Ottawa, Ont. : National Research Council of Canada; 1991 Dec.
 Canadian journal of botany; Journal canadien de botanique v.
 69 (12): p. 2806-2808; 1991 Dec.  Includes references.
 
 Language:  English
 
 Descriptors: Idaho; Rudbeckia occidentalis; Allelopathins;
 Seed germination; Germination inhibitors; Roots; Growth
 inhibitors; Lactuca sativa; Pinus contorta; Picea engelmannii;
 Plant extracts
 
 
 42                                  NAL Call. No.: TD930.A32
 Allelopathic response of vegetables to guayule residue.
 Schloman, W.W. Jr; Hilton, A.S.; McCrady, J.J.
 Essex : Elsevier Applied Science Publishers; 1991.
 Bioresource technology v. 35 (2): p. 191-196; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Parthenium argentatum; Processing; Plant
 residues; Phytotoxicity; Application to land; Seed
 germination; Germination inhibitors; Vegetables; Allelopathy;
 Water; Leachates
 
 
 43                                 NAL Call. No.: QD415.A1J6
 Allelopathic substances and interactions of Delonix regia
 (Boj) Raf. Chou, C.H.; Leu, L.L.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Dec.
 Journal of chemical ecology v. 18 (12): p. 2285-2303; 1992
 Dec.  Includes references.
 
 Language:  English
 
 Descriptors: Taiwan; Cabt; Delonix regia; Allelopathy; Leaves;
 Flowers; Plant composition; Phenolic compounds; Phytotoxicity;
 Undergrowth; Mode of action; Chemical ecology
 
 Abstract:  A unique pattern of weed exclusion was found under
 the canopy of Delonix regia, which was planted in many places
 as an ornamental tree in the south of Taiwan. A quadrat method
 was employed to examine the botanical composition between the
 area underneath D. regia and its adjacent control grassland.
 The number of species and coverage of understory species were
 significantly lower in the area of the D. regia than that of
 the grassland, indicating the growth of understory species was
 suppressed by D. regia. A series of aqueous extracts of
 leaves, flowers, and twigs of D. regia were bioassayed against
 three species to determine their phytotoxicity, and the
 results showed highest inhibition in the flowers. A water-
 culture experiment indicated the aqueous extract of flowers of
 D. regia on two local understory species (Isachne nipponensis
 and Centella asiatica) inhibited growth of both species by
 more than 70%. The phytotoxicities of fallen leaves and
 flowers of D. regia were not significantly affected by
 temperature. When the plant material was subjected to
 temperatures above 70 degrees C, however, phytotoxicity was
 decreased, indicating that the allelopathic nature of D. regia
 could easily be decomposed by fire. By means of paper, thin-
 layer, and high-performance liquid chromatography, and UV-
 visible spectrophotometry, responsible phytotoxins present in
 leaves, flowers, and twigs of D. regia were identified as 4-
 hydroxybenzoic, chlorogenic, 3,4-dihydroxybenzoic, gallic,
 3,4-dihydroxycinnamic, 3,5-dinitrobenzoic, and L-azetidine-2-
 carboxylic acids, and 3,4-dihydroxybenzaldehyde. The findings
 of bioassays and the number and amount of responsible
 allelopathic compounds found in D. regia are well correlated,
 thus permitting the conclusion that the exclusion of
 understory plants under the canopy of D. regia trees was due
 primarily to the allelopathic effect of the fallen flower,
 leaves, and twigs of the D. regia. A possible mechanism of
 action is discussed.
 
 
 44                               NAL Call. No.: Fiche no.308
 Allelopathische Effekte der Salicylsaure am Modellbeispiel von
 Vicia faba L vorgelegt von Barbara Manthe  [Allelopathic
 effects of salicyclic acid on model examples of Vicia faba L].
 Manthe, Barbara,
 1991; 1991.
 92 leaves : ill.  Vita.  Includes bibliographical references
 (leaves 85-92).
 
 Language:  German
 
 
 45                       NAL Call. No.: SB617.45.W47N69 1991
 Allelopathy.
 Stevens, K.L.
 Boulder : Westview Press; 1991.
 Noxious range weeds / edited by Lynn F. James ... [et al.]..
 p. 127-137; 1991. (Westview special studies in agriculture
 science and policy).  Includes references.
 
 Language:  English
 
 Descriptors: Weeds; Rangelands; Plant interaction; Plant
 competition
 
 
 46                                  NAL Call. No.: QH506.U34
 Allelopathy: a viable weed control strategy.
 Putnam, A.R.; Nair, M.G.; Barnes, J.P.
 New York, N.Y. : Wiley-Liss, Inc; 1990.
 UCLA symposia on molecular and cellular biology v. 112: p.
 317-322; 1990.  In the series analytic: New directions in
 biological control: Alternatives for suppressing agricultural
 pests and diseases / edited by R.R. Baker and P.E. Dunn.
 Proceedings of a UCLA Colloquium, January 20-27, 1989, Frisco,
 Colorado.  Includes references.
 
 Language:  English
 
 Descriptors: Weeds; Weed control; Allelopathy; Crops; Crop
 residues; Herbicidal properties; Plant competition
 
 Abstract:  Allelopathy is the interference plants impose upon
 one another through release of chemicals. It has been
 implicated most frequently with aggressive weeds in their
 interference with crops and less frequently with crops against
 weeds. Work in our laboratory has focused on the use of
 allelopathic crops or their residues for weed control.
 Screening of crop germplasm indicates that differential
 allelopathic potential exists within these collections. The
 most successful approach we have employed is to use
 allelopathic cereal grains in rotation with annual crops or in
 companion planting with perennial crops. Rye (Secale cereale
 L.) is an example of a plant which provides excellent weed
 suppression through both allelopathic and competitive
 mechanisms. Rye residues maintained on the soil surface
 release 2,4-dihydroxy-1,4(2H)-benzoxazin-3-one (DIBOA) and a
 breakdown product 2(3H)-benzoxazalinone (BOA) both of which
 are strongly inhibitory to germination and seedling growth of
 dicoytylenous annual weeds. In addition, soil fungi convert
 BOA to 2,2'-oxo-1,1'-azo-benzene which is ten-fold more
 phytotoxic than BOA. Hence a variety of natural products
 contribute to the herbicidal activity of rye residues.
 
 
 47                                  NAL Call. No.: 64.8 C883
 Allelopathy and autotoxicity in alfalfa: characterization and
 effects of preceding crops and residue incorporation.
 Hegde, R.S.; Miller, D.A.
 Madison, Wis. : Crop Science Society of America; 1990 Nov.
 Crop science v. 30 (6): p. 1255-1259; 1990 Nov.  Includes
 references.
 
 Language:  English
 
 Descriptors: Illinois; Medicago sativa; Sorghum bicolor;
 Rotations; Sequential cropping; Allelopathy; Allelopathins;
 Phytotoxicity; Crop residues; Roots; Shoots; Incorporation;
 Leachates; Bioassays; Seed germination; Growth rate
 
 Abstract:  Alfalfa (Medicago sativa L.) is known to be both
 autotoxic and allelopathic. Greenhouse and laboratory
 experiments were conducted to determine if 'WL-316' alfalfa
 exhibits short-term autotoxicity and long-term autotoxicity
 and allelopathy. Long-term autotoxicity and allelopathy of
 alfalfa were verified at Urbana, IL, by comparing the
 germination and growth of alfalfa and sorghum [Sorghum bicolor
 (L.) Moench] on Flanagan silt loam (fine, montmorillonitic,
 mesic Aquic Argiudoll) previously cropped to alfalfa (alfalfa-
 soil) and sorghum (sorghum-soil). Short-term autotoxicity of
 alfalfa was investigated by studying the effect of
 incorporating its roots only and both roots and shoots on the
 germination and growth of alfalfa in alfalfa-soil and sorghum-
 soil. The data were further supported by a laboratory bioassay
 of seedling exudate and shoot leachate of alfalfa and sorghum.
 Plant height and fresh weight per plant of alfalfa and fresh
 weight per plant of sorghum were lower on alfalfa-soil than on
 sorghum-soil. Germination percentages of both alfalfa and
 sorghum and plant height of sorghum were unaffected by the
 preceding crop. The two soils differed in nutrient content,
 but fertility was high and should not have been limiting to
 the growth of either crop. As a result, allelopathic/autotoxic
 compounds in alfalfa-soil were implicated in the growth
 inhibition of the two crops. Soil incorporation of fresh
 alfalfa roots only or both roots and shoots reduced alfalfa
 emergence, plant height, and dry weight per plant. Primary
 effects of water-soluble inhibitory compounds from alfalfa
 shoot appeared to be on germination and radicle elongation,
 the latter being apparently more sensitive than the former.
 Alfalfa allelopathy seems to be more severe than autotoxicity.
 A flow diagram describes different kinds of allelopathy and
 autotoxicity and various situations that verify the existence
 of a particular kind of allelopathy or autotoxicity.
 
 
 48                                 NAL Call. No.: SB611.5.S3
 Allelopathy application for control of some weed species final
 report 1985-1989.
 Saric, Taib
 Sarajevo : Faculty of Agriculture, 1990; 1990.
 31 leaves, [5] leaves of plates : ill. ; 28 cm.  Cover title. 
 Project: (USDA) JF511-11.  January 1990.  Includes
 bibliographical references (leaves 30-31).
 
 Language:  English
 
 Descriptors: Allelopathy; Weeds; Allelopathic agents
 
 
 49                                     NAL Call. No.: S51.E2
 Allelopathy as a factor in the pasture ecosystem.
 Smith, A.E.
 Athens, Ga. : The Stations; 1991 May.
 Research bulletin - University of Georgia, Agricultural
 Experiment Stations (399): 11 p.; 1991 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Georgia; Pastures; Fodder crops; Weeds;
 Allelopathy
 
 
 50                             NAL Call. No.: QK911.A46 1991
 Allelopathy basic and applied aspects.
 Rizvi, S. J. H.,_1955-; Rizvi, V.,
 New York : Chapman and Hall, 1991; 1991.
 xx, 480 p. : ill. ; 24 cm.  Includes bibliographical
 references and index.
 
 Language:  English
 
 Descriptors: Allelopathy; Allelopathic agents
 
 
 51                                  NAL Call. No.: SD112.F67
 Allelopathy in barley: potential for biological suppression of
 weeds. Liu, D.L.; Lovett, J.V.
 Rotorua : The Institute; 1990.
 FRI bulletin - Forest Research Institute, New Zealand Forest
 Service (155): p. 85-92. ill; 1990.  Paper presented at the
 "Conference on Alternatives to the Chemical Control of Weeds,"
 held July 25-27, 1989, Rotorua, New Zealand. Includes
 references.
 
 Language:  English
 
 Descriptors: Hordeum vulgare; Allelopathy; Seed germination;
 Radicles; Sinapis alba; Allelochemicals; Gramine; Hordenine;
 Phytotoxicity; Biological control; Weed control
 
 
 52                                    NAL Call. No.: 4 AM34P
 Allelopathy of crop residues influences corn seed germination
 and early growth.
 Martin, V.L.; McCoy, E.L.; Dick, W.A.
 Madison, Wis. : American Society of Agronomy; 1990 May.
 Agronomy journal v. 82 (3): p. 555-560; 1990 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Ohio; Zea mays; Seed germination; Growth;
 Inhibition; Allelopathy; Crop residues; Oats; Soy straw;
 Soybeans; Microbial activities; Phytotoxicity; Temperature;
 Aeration; Decomposition
 
 Abstract:  Crop residues produce alleochemicals that may
 inhibit corn [Zea mays (L.)] seed germination and early
 growth. Studies were conducted in which residues of corn,
 soybean [Glycine max (L.) Merr.], oat [Avena sativa (L.)], and
 mixed grass hay were extracted under N2 gas or air. Organic
 debris was removed and half of each extract was filter
 sterilized. Corn seeds were incubated in the extracts for 96 h
 at 25 degrees C. Percent germination, and lengths of
 coleoptile, radicle, and secondary roots were measured.
 Residues extracted under N2 gas or air did not differ
 significantly in their toxicity. Nonsterile residue extracts
 decreased germination to 74% for soybean and oat straw and 27%
 for corn and hay residues. Sterile extracts did not affect
 germination. Nonsterile soybean and oat extracts did not
 reduce coleoptile lengths but did reduce radicle and secondary
 root lengths by 34% compared to the water treatment.
 Sterilized extracts reduced radicle and secondary root lengths
 by 63%. Nonsterile corn and hay extracts reduced coleoptile
 lengths by 42% and radicle and secondary root lengths by 81%.
 A second extraction was performed by incubating the residues
 without aeration at 25 and 0.5 degrees C. Seed germination for
 treatments with nonsterile extracts obtained at 25 degrees C
 were similar to those for nonsterile extracts of Exp. 1.
 Extraction at 0.5 degrees C and filter sterilization also
 improved germination. Soybean and oat extracts did not
 strongly inhibit coleoptile lengths; however, a 61% reduction
 occurred in radicle and secondary root lengths for the
 sterilized, 0.5 degrees C extract. Corn and hay residues were
 generally more inhibitory to coleoptile, radicle and secondary
 root lengths; however, no consistent effects were observed
 from temperature and sterilization treatments.
 
 
 53                                 NAL Call. No.: QD415.A1J6
 Allelopathy of Sasa cernua.
 Li, H.H.; Nishimura, H.; Hasegawa, K.; Mizutani, J.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Oct.
 Journal of chemical ecology v. 18 (10): p. 1785-1796; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Sasa; Allelopathy; Volatile compounds; Phenolic
 compounds; Growth inhibitors; Allelochemicals; Rhizosphere;
 Weed control
 
 Abstract:  Sasa (Sasa cernua Makino) is a very serious weed
 pest. Its allelopathy was studied using lettuce, wheat,
 timothy, and green amaranth as testing species, Cultured in
 the rhizosphere soil of Sasa cernua, the seedlings were
 inhibited by 42-80% compared with the controls cultured in
 normal soil and vermiculite. The phenolic fraction extracted
 with 1 M NaOH from the rhizosphere soil of S. cernua caused
 significant inhibitions on the seed germination seedling
 growth of lettuce, timothy, green amaranth, and barnyard
 grass. p-Coumaric, ferulic, vanillic, and p-hydroxybenzoic
 acids and p-hydroxybenzaldehyde were identified as the main
 allelochemicals in sasa soil by HPLC and [1H]NMR. Their
 contents in the rhizosphere soil were 5640, 1060, 860, 810 and
 630 micrograms/100 g soil. The neutral fraction inhibited the
 seed germination and seedling growth of lettuce in the TLC
 direct bioassay. Volatile compounds released from sasa leaves
 also inhibited the growth of lettuce, wheat, timothy, and
 green amaranth grown under light, and the growth of etiolated
 seedlings of barley and wheat. These results confirm that S.
 cernua produces typical allelopathy through its rhizosphere
 soil and air space.
 
 
 54                                 NAL Call. No.: QD415.A1J6
 Allelopathy of small everlasting (Antennaria microphylla)
 phytotoxicity to leafy spurge (Euphorbia esula) in tissue
 culture.
 Hogan, M.E.; Manners, G.D.
 New York, N.Y. : Plenum Press; 1990 Mar.
 Journal of chemical ecology v. 16 (3): p. 931-939; 1990 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Antennaria microphylla; Extracts; Callus; Cell
 suspensions; Phytotoxicity; Euphorbia esula
 
 Abstract:  Media and media extracts from callus cultures of
 small everlasting (Antennaria microphylla) inhibited leafy
 spurge (Euphorbia esula L.) callus tissue and suspension
 culture growth (50 and 70% of control, respectively) and were
 phytotoxic in lettuce and leafy spurge root elongation
 bioassays (64 and 77% of control, respectively). Hydroquinone,
 a phytotoxic compound previously isolated from small
 everlasting, was also biosynthesized by callus and suspension
 cultures of this species. Exogenously supplied hydroquinone
 (0.5 mM) was toxic to leafy spurge suspension culture cells
 and was only partially biotransformed to its nontoxic water-
 soluble monoglucoside, arbutin, by these cells. This report
 confirms the chronic involvement of hydroquinone in the
 allelopathic interaction between small everlasting and leafy
 spurge.
 
 
 55                                 NAL Call. No.: QD415.A1J6
 Allelopathy of yellow fieldcress (Rorippa sylvestris):
 identification and characterization of phytotoxic
 constituents.
 Yamane, A.; Nishimura, H.; Mizutani, J.
 New York, N.Y. : Plenum Press; 1992 May.
 Journal of chemical ecology v. 18 (5): p. 683-691; 1992 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Rorippa sylvestris; Allelopathy; Root exudates;
 Plant composition; Bioassays; Seed germination; Inhibition;
 Seedlings; Growth; Lactuca sativa; Weed control
 
 Abstract:  Both the neutral and acidic fractions of the
 acetone extract of yellow fieldcress (Kireha-inugarashi,
 Rorippa sylvestris Besser) inhibited lettuce seed germination.
 Salicylic, p-hydroxybenzoic, vanillic, and syringic acid were
 identified in the acidic fraction. In the neutral fraction,
 hirsutin (8-methylsulfinyloctyl isothiocyanate), 4-
 methoxyindole-3-acetonitrile, and pyrocatechol were
 identified. Bioassay using a root exudate recirculating system
 showed R. sylvestris during flowering inhibited the lettuce
 seedling growth. Hirsutin (13 micrograms/plant/day) and
 pyrocatechol (9.3 micrograms/plant/day) were the major
 compounds released into the rhizosphere. Several combinations
 of pyrocatechol, p-hydroxybenzoic acid, vanillic acid, and
 hirsutin reduced lettuce seedling growth. These compounds
 seemed to be allelochemicals.
 
 
 56                                  NAL Call. No.: aZ5071.N3
 Allelopathy: the effects of chemicals produced by plants,
 January 1986-January 1990.
 Gilbert, H.
 Beltsville, Md. : The Library; 1990 Apr.
 Quick bibliography series - U.S. Department of Agriculure,
 National Agricultural Library (U.S.). (90-46): 28 p.; 1990
 Apr.  Updates QB 88-62. Bibliography.
 
 Language:  English
 
 Descriptors: Plants; Allelopathy; Allelopathins;
 Phytotoxicity; Chemical constituents of plants; Bibliographies
 
 
 57                                  NAL Call. No.: aZ5071.N3
 Allelopathy: the effects of chemicals produced by plants--
 January 1988-April 1992.
 Gilbert, H.
 Beltsville, Md. : The Library; 1992 Jun.
 Quick bibliography series - U.S. Department of Agriculture,
 National Agricultural Library (U.S.). (92-50): 65 p.; 1992
 Jun.  Updates QB 90-46. Bibliography.
 
 Language:  English
 
 Descriptors: Plants; Allelopathy; Allelopathins;
 Phytotoxicity; Bibliographies
 
 
 58                                  NAL Call. No.: QK911.A44
 Allelopatiia i produktivnost' rastenii sbornik nauchnykh
 trudov  [Allelopathy and the productivity of plants].
 Grodzinskii, A. M.
 TSentral
 Kiev : Nauk. dumka, 1990; 1990.
 146 p. : ill. ; 20 cm.  "Nauchnoe izdanie"--Colophon.  At head
 of title: Akademiia nauk Ukrainskoi SSR. TSentral.  Includes
 bibliographical references.
 
 Language:  Russian
 
 Descriptors: Allelopathy; Allelopathic agents; Plant
 physiology
 
 
 59                           NAL Call. No.: QK898.A43M6 1990
 Allelopatiia v plodovykh sadakh  [Allelopathy in orchards].
 Moroz, P. A.
 Kiev : Haukova dumka, 1990; 1990.
 208 p., [4] p. of plates : ill. ; 21 cm.  At head of title:
 Akademiia nauk Ukrainskoi SSR. Tsentral'nyi respublikanskii
 botanicheskii sad.
 
 Language:  Russian
 
 Descriptors: Fruit trees; Allelopathic agents
 
 
 60                                    NAL Call. No.: 81 SO12
 Allelpathic potential of celery residues on lettuce.
 Shilling, D.G.; Dusky, J.A.; Mossler, M.A.; Bewick, T.A.
 Alexandria, Va. : The Society; 1992 Mar.
 Journal of the American Society for Horticultural Science v.
 117 (2): p. 308-312; 1992 Mar.  Includes references.
 
 Language:  English
 
 Descriptors: Lactuca sativa; Seedlings; Plant residues; Soil;
 Incorporation; Apium graveolens; Allelopathy; Phytotoxins;
 Seedling emergence; Growth; Adverse effects; Activated carbon;
 Greenhouse culture
 
 Abstract:  Poor emergence of commercially grown lettuce has
 been observed when planted immediately after the removal of a
 celery crop. Greenhouse experiments were conducted to evaluate
 the possible allelopathic effects of celery residue on the
 emergence and growth of lettuce. The influence of amount and
 type of celery tissue, growth medium and fertility, incubation
 time in soil, and amendment of growth medium containing celery
 residue with activated charcoal was evaluated with respect to
 the allelopathic potential of celery. Celery root tissue was
 1.8 and 1.6 times more toxic to lettuce seedling growth than
 was celery petiole or lamina tissue, respectively. Lettuce
 shoot growth was inhibited to a greater extent when grown in
 sand amended with celery residue rather than either amended
 vermiculite or potting soil. Incubation of celery root residue
 in soil for 4 weeks increased phytotoxicity at 1% (v/v) and
 decreased it at 40% (v/v). Increasing the fertility of pure
 sand with varying amounts of Hoagland's solution did not
 reverse the allelopathic effects of celery residue. The
 addition of activated carbon to the medium increased the
 growth of lettuce exposed to celery residues. Celery residues
 possess allelopathic potential to developing lettuce
 seedlings. Celery tissue type and concentration, soil type,
 incubation of celery root residue in soil, and addition of
 activated carbon to the growing medium influenced the
 magnitude of the observed phytotoxicity.
 
 
 61                                   NAL Call. No.: 450 AM36
 Allyl isothiocyanate release and the allelopathic potential of
 Brassica napus (Brassicaceae).
 Choesin, D.N.; Boerner, R.E.J.
 Columbus, Ohio : Botanical Society of America; 1991 Aug.
 American journal of botany v. 78 (8): p. 1083-1090; 1991 Aug. 
 Includes references.
 
 Language:  English
 
 Descriptors: Brassica napus; Allyl isothiocyanate;
 Biosynthesis; Allelopathins; Plant interaction; Growth
 inhibitors; Plant density; Mutants; Genotypes; Genetic
 variation
 
 Abstract:  The allelopathic potential of Brassica species has
 been attributed to release of the mustard oil glycosides which
 they produce in large quantities. Upon hydrolysis, these
 glucosinolates yield isothiocyanates, compounds with strong
 antibiotic properties. The objective of this study was to
 assess whether Brassica napus, a common and widespread crop
 and weed crucifer, is capable of allelopathic interference,
 and if so, whether its glycoside derivative, allyl
 isothiocyanate (AI), is capable of producing this
 interference. Wild type and low glucosinolate-mutant B. napus
 were grown in low organic content soil under greenhouse
 conditions, and AI release into soil was monitored. Most
 plants released low levels of AI, though approximately 10%
 released much higher levels. Wild type plants released more AI
 than mutants. Growth of the target species, Medicago sativa,
 was not affected by additions of AI to soils at concentrations
 equal to the median and 95% quantile from the B. napus soils.
 In replacement series experiments, the two B. napus genotypes
 suppressed growth of M. sativa equally despite differences in
 AI release rate. In an intraspecific replacement series
 experiment, the two B. napus genotypes were equal competitors.
 Under our experimental conditions, B. napus showed no
 indication of being allelopathic, and AI concentrations
 typical of soils around B. napus plants did not inhibit target
 plants.
 
 
 62                                    NAL Call. No.: QK1.A28
 Antibiotic effect of Rhizobium sp. towards some soil fungi.
 Anbu, D.A.; Sullia, S.B.
 Meerut, India : Society for Advancement of Botany; 1990 Dec.
 Acta botanica Indica v. 18 (2): p. 213-215; 1990 Dec. 
 Includes references.
 
 Language:  English
 
 Descriptors: India; Australia; Arachis hypogaea; Rhizobium;
 Geographical races; Plant extracts; Antibiotics; Antifungal
 agents; Rhizosphere fungi; Allelopathins; Strain differences
 
 
 63                                    NAL Call. No.: 81 SO12
 Asparagus emergence in Fusarium-treated and sterile media
 following exposure of seeds or radicles to one or more
 cinnamic acids.
 Peirce, L.C.; Miller, H.G.
 Alexandria, Va. : The Society; 1993 Jan.
 Journal of the American Society for Horticultural Science v.
 118 (1): p. 23-28. ill; 1993 Jan.  Includes references.
 
 Language:  English
 
 Descriptors: Asparagus officinalis; Allelopathy; Cinnamic
 acid; Emergence; Growth inhibitors; Radicles; Seeds; Toxicity;
 Crop damage; Fusarium
 
 Abstract:  Several cinnamic acids have been identified as
 principal toxic components of asparagus (Asparagus officinalis
 L.) root autotoxin and have been shown to synergize Fusarium
 infection of asparagus. The basis for this synergism was
 studied by exposing asparagus seeds and radicles from
 pregerminated seeds to ferulic (FA), caffeic (CA), or
 methylenedioxycinnamic (MDA) acids alone and in combinations
 of two or three of these acids. After treatment, seeds were
 placed in pots of peat-lite mix, and, depending on the
 experiment, all or half were inoculated with F. oxysporum
 (Schlecht) f. sp. asparagi (Cohen). Seedling emergence from
 each pot was used as a measure of toxicity. All cinnamic acids
 at 1% suppressed emergence compared with the control.
 Solutions combining FA and CA (0.5%/0.5%, v/v) were
 substantially more toxic than 1% solutions of either alone.
 Exposure of radicles (early postgermination) for 10 minutes to
 combined FA/CA before planting decreased emergence from pots,
 whereas emergence following a 10-minute exposure to 1% CA or
 FA alone did not differ from the controls. The 2-hour exposure
 to FA or to FA/CA and the 24-hour exposure to CA, FA, or FA/CA
 decreased emergence, with toxicity progressing as follows: CA
 < FA < FA/CA. Root tip squashes showed fewer mitotic figures
 in treated than in untreated radicles, and scanning electron
 microscopic (SEM) examination of the radicle epidermis
 revealed damage to the surface of epidermal cells and
 precocious root hair development, the extent of which
 paralleled treatment toxicity.
 
 
 64                                 NAL Call. No.: QD415.A1J6
 Assessment of allelopathic potential in Artemisia princeps
 var. orientalis residues.
 Yun, K.W.; Kil, B.S.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Nov.
 Journal of chemical ecology v. 18 (11): p. 1933-1940; 1992
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Artemisia princeps; Plant composition;
 Allelopathy; Leaves; Plant residues; Phytotoxicity; Bioassays;
 Growth inhibitors; Seedlings
 
 Abstract:  Field and laboratory studies were conducted to
 examine the differential phytotoxicity of residues of
 Artemisia princeps var. orientalis (wormwood) using various
 plants as test species. Seedling elongation and dry weights of
 receptor plants were inversely proportional to the
 concentration and incubation time of dry leaves of A. princeps
 var. orientalis in vermiculite. In seedling growth tests with
 abandoned field soils (control) and soil underneath wormwood
 plants (test), the elongation, dry weight, and caloric content
 of seedlings grown in the soil from under wormwood plants were
 severely inhibited, thereby suggesting that certain growth
 inhibitors were released from wormwood and the inhibitor
 remained in the soil.
 
 
 65                                   NAL Call. No.: 450 AU72
 An assessment of the allelopathic potential of Eucalyptus.
 May, F.E.; Ash, J.E.
 East Melbourne : Commonwealth Scientific and Industrial
 Research Organization; 1990.
 Australian journal of botany v. 38 (3): p. 245-254; 1990. 
 Includes references.
 
 Language:  English
 
 Descriptors: Australian capital territory; Eucalyptus
 globulus; Eucalyptus maculata; Eucalyptus macrorhyncha;
 Eucalyptus rossii; Eucalyptus rubida; Allelopathins;
 Leachates; Bark; Leaves; Forest litter; Stemflow; Laboratory
 methods; Allelopathy; Bioassays
 
 
 66                                  NAL Call. No.: 421 EN895
 Azadirachtin inhibits secretion of trypsin in midgut of
 Manduca sexta caterpillars: reduced growth due to impaired
 protein digestion. Timmins, W.A.; Reynolds, S.E.
 Dordrecht : Kluwer Academic Publishers; 1992 Apr.
 Entomologia experimentalis et applicata v. 63 (1): p. 47-54;
 1992 Apr. Includes references.
 
 Language:  English
 
 Descriptors: Manduca sexta; Midgut; Secretion; Trypsin;
 Azadirachtin; Allelochemicals; Antifeedants; Growth
 inhibitors; Protein digestion; Proteinases
 
 
 67                                   NAL Call. No.: SB925.B5
 Behavioral and ecological constraints imposed by plants on
 insect parasitoids: implications for biological control.
 Kester, K.M.; Barbosa, P.
 Orlando, Fla. : Academic Press; 1991 Aug.
 Biological control v. 1 (2): p. 94-106; 1991 Aug.  Paper
 presented at the "Symposium on Host/Parasitoid Interactions,"
 December 3, 1990, New Orleans, Louisiana.  Includes
 references.
 
 Language:  English
 
 Descriptors: Plant pests; Cotesia; Manduca sexta; Biological
 control; Parasites of insect pests; Allelochemicals;
 Adaptation; Nicotine; Trophic levels; Host parasite
 relationships; Feeding behavior; Plant composition
 
 
 68                                  NAL Call. No.: 450 P5622
 beta-(3-isoxazolin-5-on-2-yl)-alanine from Pisum: allelopathic
 properties and antimycotic bioassay.
 Schenk, S.U.; Werner, D.
 Oxford : Pergamon Press; 1991.
 Phytochemistry v. 30 (2): p. 467-470; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Pisum sativum; Seedlings; Root exudates; Chemical
 analysis; Alanine; Derivatives; Allelopathins; Gramineae;
 Lactuca sativa; Germinationinhibitors; Growth inhibitors;
 Antifungal properties
 
 Abstract:  Grasses and Lactuca sativa when germinated in the
 presence of the non-protein amino acid beta-(3-isoxazolin-5-
 on-2-yl)-alanine (betaIA) from roots and root exudates of pea
 seedlings, showed a pronounced reduction of root length and a
 necrosis of the root tips. Growth of legume seedlings was only
 slightly affected. We suggest the role of this secondary plant
 product as an allelochemical. Besides its effect on plant
 morphogenesis, betaIA also exhibits an antimycotic activity
 towards Saccharomyces cerevisiae with a minimum inhibitory
 concentration (MIC) of 0.5 ppm.
 
 
 69                                   NAL Call. No.: QL495.A7
 The biochemical and physiological effects of insect hosts on
 the development and ecology of their insect parasites: an
 overview.
 Lawrence, P.O.
 New York, N.Y. : Wiley-Liss; 1990.
 Archives of insect biochemistry and physiology v. 13 (3/4): p.
 217-228; 1990. Literature review.  Includes references.
 
 Language:  English
 
 Descriptors: Insect pests; Parasites of insect pests; Host
 parasite relationships; Allelochemicals; Molting hormones;
 Diapause; Metamorphosis
 
 
 70                                  NAL Call. No.: 450 P5622
 Biochemical basis for the resistance of barley to aphids.
 Corcuera, L.J.
 Oxford ; New York : Pergamon Press, 1961-; 1993 Jul.
 Phytochemistry v. 33 (4): p. 741-747; 1993 Jul.  Includes
 references.
 
 Language:  English
 
 Descriptors: Hordeum; Diuraphis; Metopolophium; Rhopalosiphum;
 Schizaphis; Sitobion; Insect pests; Pest resistance;
 Allelochemicals; Defense mechanisms; Induction; Plant
 morphology; Barriers; Environmental factors; Plant nutrition;
 Literature reviews
 
 Abstract:  Barley plants may be severely damaged by aphids,
 mainly because they may transmit viruses, remove essential
 nutrients, and disrupt tissues. This review discusses possible
 resistance factors of this plant, such as morphological
 defences and natural chemicals that have been shown or
 suggested to be involved in protection of barley against
 aphids. The available evidence for the role played by waxes,
 gramine, aconitic acid, phenolics and amino acids is
 presented. A discussion is included on other potentially
 protective molecules, such as protease inhibitors, that need
 to be studied. Environmental stress also affects plant-aphid
 interactions because the chemical composition of the plant
 changes. Water stress increases susceptibility, and Nacl and
 temperature increase resistance to aphids. The compatible
 solute glycine betaine, which accumulates under several types
 of stress. increases reproduction of aphids. Temperature and
 availability of nitrates increase gramine content of the
 leaves and. therefore, resistance to the aphids. A summary of
 conclusions and future perspectives focuses on the paramount
 importance of environmental stress in plant resistance and on
 the need to identify inducible resistance factors.
 
 
 71                                 NAL Call. No.: QD415.A1B5
 Biochemical defence of pro-oxidant plant allelochemicals by
 herbivorous insects.
 Ahmad, S.
 Oxford ; New York : Pergamon Press, 1974-; 1992 Jun.
 Biochemical systematics and ecology v. 20 (4): p. 269-296;
 1992 Jun.  Includes references.
 
 Language:  English
 
 Descriptors: Plant composition; Oxygen; Allelochemicals;
 Antioxidants; Papilio polyxenes; Spodoptera eridania;
 Trichoplusia ni; Metabolic detoxification; Literature reviews;
 Chemical ecology
 
 Abstract:  A new aspect of interactions among insect
 herbivores and defensive chemistry of plants in the regulation
 of oxygen toxicity exerted by pro-oxidant allelochemic is
 described. Endogenous oxygen toxicity results from activation
 of the ground state of molecular oxygen to the superoxide
 anion radical (O2.-), hydrogen peroxide (H2O2), hydroxyl
 radical (.OH), lipid hydroperoxides (LOOHs), and peroxyl
 radicals (LO2. or RO2.). The strongly lipid-peroxidizing
 singlet oxygen (1 delta g O2) is also produced during light
 activation of photosensitizers. Ingestion of pro-oxidants
 exacerbates oxygen toxicity by increasing the production of
 these deleterious forms of oxygen. The role of ascorbate,
 alpha-tocopherol, glutathione, carotenoids and urate as
 antioxidants in insects is apparent, but needs more work for
 the elucidation of their roles. The major defence mechanism
 includes a group of antioxidant enzymes represented by
 superoxide dismutase (SOD), catalase (CAT), glutathione-S-
 transferase's peroxidative activity (GSTPX), glutathione
 reductase (GR), and DT-diaphorase. SOD converts O2.- radicals
 to H2O2 and 2, CAT decomposes H2O2 to H2O and O2, GSTPX
 reduces LOOHs to LOHs with GSH as reductant, and GSSG formed
 from GSH during the GSTPX reaction is reduced to GSH by GR.
 DT-diaphorase is an important antioxidant in that it reduces
 quinones by a two-electron reduction to stable products,
 thereby preventing the one-electron reduction to semiquinone
 radicals which generate O2 radicals. Therefore, these enzymes
 are crucial for insect herbivores for preventing the free-
 radical cascade of oxygen, and terminating the toxic lipid
 peroxidation chain reaction, in response to the endogenous and
 potential exogenous oxidant-induced injury.
 
 
 72                                   NAL Call. No.: QL495.A7
 Bioengieering of crop plants and resistant biotype evolution
 in insects: counteracting coevolution.
 Brattsten, L.B.
 New York, N.Y. : Wiley-Liss; 1991.
 Archives of insect biochemistry and physiology v. 17 (4): p.
 253-267; 1991. Paper presented at a symposium on biochemical
 strategies of offense and defense at the plant-insect
 interface, 1989, San Antonio, Texas.  Literature review. 
 Includes references.
 
 Language:  English
 
 Descriptors: Insect pests; Pest resistance; Genetic
 engineering; Allelochemicals; Insecticide resistance;
 Literature reviews
 
 Abstract:  The use, as opposed to the procurement, of
 transgenic crop plants is discussed in this paper. Transgenic
 crop plants must not be used until appropriate strategies for
 their use have been designed and not before crop plants with a
 variety of insect defenses have been developed. The use of a
 crop plant with a single defense will pose as strong a
 selection pressure as the use of a single synthetic
 insecticide, since insect herbivores are able to evolve
 effective counter-defenses. The defenses of insects in natural
 plant-insect associations and with regard to synthetic
 insecticides are described to demonstrate that there is
 nothing unique about insecticide resistance. It is the
 inevitable alternative to local extinction in response to a
 persistent and predictable selection pressure. Plants
 counteract insect defensive evolution by keeping the selection
 pressure as variable as possible. This leads to the conclusion
 that the best use of biotechnology in crop protection is to
 reintroduce chemical diversity into crop plants.
 
 
 73                                 NAL Call. No.: SB950.A1P3
 Biological control of Parthenium hysterophorus L. (Asteraceae)
 by Cassia uniflora Mill (Leguminosae), in Bangalore, India.
 Joshi, S.
 London : Taylor & Francis; 1991 Apr.
 Tropical pest management v. 37 (2): p. 182-184; 1991 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Karnataka; Parthenium hysterophorus; Weed
 control; Biological control; Cassia; Biological control
 agents; Competitive ability; Allelopathy; Seeds; Leachates;
 Germination inhibitors; Seed germination
 
 
 74                                  NAL Call. No.: 450 P5622
 Biologically active labdane-type diterpene glycosides from the
 root-stalks of Gleichenia japonica.
 Munesada, K.; Siddiqui, H.L.; Suga, T.
 Oxford : Pergamon Press; 1992 May.
 Phytochemistry v. 31 (5): p. 1533-1536; 1992 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Japan; Gleichenia japonica; Roots; Chemical
 composition; Diterpenes; Glycosides; Growth inhibitors;
 Allelopathins; Lactuca sativa
 
 Abstract:  A glycoside showing a strong growth inhibition of
 lettuce was isolated from the root-stalks of Gleichenia
 japonica and its structure was established to be the 3-O-
 alpha-rhamnopyranosyl-(1 leads to 2)-beta-glucopyranoside of
 13-O-7-rhamnopyranosyl-(+)-3 beta-hydroxymanool. In addition,
 two related glycosides were also isolated and they were
 characterized as the 3-O-beta-fucopyranosyl-(1 leads to
 3)-alpha-rhamnopyranosyl-(1 leads to 2)-beta-glucopyranoside
 of 13-O-alpha-rhamnopyranosyl-(+)-3 beta-hydroxymanool and the
 13-O-rhamnopyranoside of the same diterpene alcohol. The
 diterpene alcohol accelerated the growth of lettuce.
 
 
 75                                  NAL Call. No.: SB610.R47
 Biology and control of morningglories (Ipomoea spp.).
 Elmore, C.D.; Hurst, H.R.; Austin, D.F.
 Champaign, Ill. : Weed Science Society of America; 1990.
 Reviews of weed science v. 5: p. 83-114. ill; 1990. 
 Literature review. Includes references.
 
 Language:  English
 
 Descriptors: Ipomoea; Weed biology; Seed germination; Sexual
 reproduction; Asexual reproduction; Taxonomy; Keys;
 Competitive ability; Allelopathy; Weed control; Perennial
 weeds; Annual habit; Chemical control; Biological control;
 Literature reviews
 
 
 76                                    NAL Call. No.: QK1.C83
 Bioregulator-induced changes in allelochemicals and their
 effects on plant resistance to pests.
 Hedin, P.A.
 Boca Raton, Fla. : CRC Press; 1990.
 Critical reviews in plant sciences v. 9 (5): p. 371-379; 1990. 
 Literature review.  Includes references.
 
 Language:  English
 
 Descriptors: Plant pests; Pest resistance; Pest control;
 Allelochemicals; Plant growth regulators; Chemical analysis;
 Insecticidal properties; Plant extracts; Literature reviews
 
 
 77                                NAL Call. No.: SB950.A1I66
 Brassica alternatives to herbicides and soil fumigants.
 Grossman, J.
 Berkeley, CA : Bio-Integral Resource Center,; 1993 Jul.
 The IPM practitioner : the newsletter of integrated pest
 management v. 15 (7): p. 1-10; 1993 Jul.  Includes references.
 
 Language:  English
 
 Descriptors: Crops; Weed control; Brassica; Biological
 control; Live mulches; Allelopathy; Cover crops; Plant disease
 control; Pest control; Cultural control; Nematode control;
 Green manures; Soil solarization; Plantparasitic nematodes;
 Alternative farming
 
 
 78                                  NAL Call. No.: 421 En895
 Caterpillars' compensatory feeding response to diluted
 nutrients leads to toxic allelochemical dose.
 Slansky, F. Jr; Wheeler, G.S.
 Dordrecht : Kluwer Academic Publishers; 1992 Nov.
 Entomologia experimentalis et applicata v. 65 (2): p. 171-186;
 1992 Nov. Includes references.
 
 Language:  English
 
 Descriptors: Camellia; Coffea; Hosts of plant pests;
 Anticarsia gemmatalis; Larvae; Plant pests; Toxicity;
 Allelochemicals; Caffeine; Eating rates; Feeding behavior;
 Nutrients
 
 
 79                                   NAL Call. No.: 421 J822
 Cell culture bioassay to evaluate allelochemical toxicity to
 Heliothis virescens (Lepidoptera: Noctuidae).
 Stipanovic, R.D.; Elissalde, M.H.; Altman, D.W.; Norman, J.O.
 Lanham, Md. : Entomological Society of America; 1990 Jun.
 Journal of economic entomology v. 83 (3): p. 737-741; 1990
 Jun.  Includes references.
 
 Language:  English
 
 Descriptors: Gossypium; Toxic exudates; Toxicity; Bioassays;
 Pest resistance; Heliothis virescens; Larvae
 
 Abstract:  An insect tissue culture bioassay was developed
 with an established cell line (BCIRL-HV-AM1) of Heliothis
 virescens (F.). This bioassay substantially reduced the time,
 material, and experimental error involved in toxicity
 evaluations compared with larval-feeding studies with
 artificial diets. LD50's of seven terpenes from the cotton
 plant (Gossypium spp.) were determined in the tissue culture
 bioassay. Various levels of toxicity were observed. Gossypol,
 hemigossypolone, and heliocides H1, H2, and H3 had LD50's of
 10 to 16 micrograms/ml. For caryophyllene oxide and
 caryophyllene, LD50's were 53 micrograms/ml and 221
 micrograms/ml, respectively. Comparison of these values with
 ED50's obtained in larval-feeding studies validate the cell
 bioassay as an effective in vitro assay, for relative
 toxicity. Twelve other terpenes of unknown toxicity to H.
 virescens also were evaluated.
 
 
 80                                  NAL Call. No.: S605.5.B5
 Changing perceptions of allelopathy and biological control.
 Lovett, J.V.
 Oxon : A B Academic Publishers; 1991.
 Biological agriculture and horticulture : an international
 journal v. 8 (2): p. 89-100; 1991.  Includes references.
 
 Language:  English
 
 Descriptors: Alternative farming; Sustainability; Farming
 systems; Biological control; Allelopathy; Allelochemicals;
 Responses; Plant protection; Weed control; Biological control
 agents; Mycoherbicides; Cost benefit analysis; Control
 methods; Crop production; Reviews
 
 
 81                                   NAL Call. No.: 451 L64J
 The chemical composition of Astragalus: a comparison of
 seleniferous and non-seleniferous plants growing side by side.
 Cowgill, U.M.; Landenberger, B.D.
 London : Academic Press; 1992 Jun.
 Botanical journal of the Linnean Society v. 109 (2): p.
 223-234; 1992 Jun. Includes references.
 
 Language:  English
 
 Descriptors: Astragalus; Species; Chemical composition;
 Selenium; Phytotoxicity; Allelochemicals; Phenolic acids;
 Flavonoids; Allelopathy; Site types
 
 
 82                                  NAL Call. No.: SD112.F67
 Chemicals in plant protection: Is there a natural
 alternative?. Lovett, J.V.
 Rotorua : The Institute; 1990.
 FRI bulletin - Forest Research Institute, New Zealand Forest
 Service (155): p. 57-65; 1990.  Paper presented at the
 "Conference on Alternatives to the Chemical Control of Weeds,"
 held July 25-27, 1989, Rotorua, New Zealand. Includes
 references.
 
 Language:  English
 
 Descriptors: Plant protection; Pesticides; Allelochemicals;
 Allelopathy; Integrated pest management
 
 
 83                                 NAL Call. No.: QD415.A1J6
 Chemotypes of Cyperus rotundus in Pacific Rim and Basin:
 distribution and inhibitory activities of their essential
 oils.
 Komai, K.; Tang, C.S.; Nishimoto, R.K.
 New York, N.Y. : Plenum Press; 1991 Jan.
 Journal of chemical ecology v. 17 (1): p. 1-8; 1991 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cyperus rotundus; Tubers; Chemical composition;
 Plant composition; Essential oils; Allelopathy;
 Sesquiterpenes; Geographical distribution
 
 Abstract:  Four major chemotypes of Cyperus rotundus L.
 (purple nutsedge) have been reported based on the composition
 of essential oils in mature tubers. Distribution of the H, M,
 K, and O type in countries of the Pacific Rim and Basin was
 investigated. In general, the H type dominates on the islands
 of Japan, and the O type has the widest range of distribution.
 The O type also dominates the Pacific Basin islands except for
 Hawaii, where the K-type is dominant. Inhibitory activity of
 the essential oils from C. rotundus tubers against the
 seedling growth of lettuce and oats was in the order of H > M
 > K > O. Seven major sesquiterpenes were isolated from the
 oils and their inhibitory activities determined. Results
 suggest that C. rotundus of different chemotypes may have
 different allelopathic activity in the crop-weed interaction.
 
 
 84                                  NAL Call. No.: SB951.P49
 Comparative metabolism of the phototoxic allelochemical alpha-
 terthienyl in three species of lepidopterans.
 Iyengar, S.; Arnason, J.T.; Philogene, B.J.R.; Werstiuk, N.H.;
 Morand, P. Duluth, Minn. : Academic Press; 1990 Jun.
 Pesticide biochemistry and physiology v. 37 (2): p. 154-164;
 1990 Jun. Includes references.
 
 Language:  English
 
 Descriptors: Manduca sexta; Heliothis virescens; Ostrinia
 nubilalis; Metabolism; Larvae; Allelopathins; Compositae;
 Thiophene; Enzyme activity; Metabolites; Oxidoreductases
 
 
 85                                   NAL Call. No.: QL495.A7
 Comparative processing of allelochemicals in the Papilionidae
 (Lepidoptera). Berenbaum, M.R.
 New York, N.Y. : Wiley-Liss; 1991.
 Archives of insect biochemistry and physiology v. 17 (4): p.
 213-221; 1991. Paper presented at a symposium on biochemical
 strategies of offense and defense at the plant-insect
 interface, 1989, San Antonio, Texas.  Includes references.
 
 Language:  English
 
 Descriptors: Papilionidae; Allelochemicals; Detoxification;
 Cytochrome p-450; Coumarins
 
 Abstract:  Within the family Papilionidae (Lepidoptera),
 species display a broad range of feeding patterns, from
 oligophagy on a single hostplant family to polyphagy on over a
 dozen families. Accompanying this diversity of feeding
 strategies is a diversity of physiological mechanisms for
 processing hostplant allelochemicals. Studies on members of
 this family as well as other Lepidoptera suggest that
 oligophagy is associated with high activity, in addition to
 high substrate specificity, of detoxicative enzymes.
 
 
 86                                  NAL Call. No.: 381 J8223
 Comparative study of proteinase inhibitors in tropical root
 crops and survey of allelochemicals in the edible aroids.
 Bradbury, J.H.; Hammer, B.C.
 Washington, D.C. : American Chemical Society; 1990 Jul.
 Journal of agricultural and food chemistry v. 38 (7): p.
 1448-1453; 1990 Jul. Includes references.
 
 Language:  English
 
 Descriptors: Alocasia macrorrhiza; Ipomoea batatas;
 Cyrtosperma chamissonis; Dioscorea alata; Dioscorea esculenta;
 Xanthosoma sagittifolium; Protease inhibitors; Trypsin
 inhibitors; Taro; Sweet potatoes; Yams; Resistance to
 parasites
 
 
 87                                  NAL Call. No.: 450 B6527
 Competition and allelopathy in aquatic plant communities.
 Gopal, B.; Goel, U.
 Bronx, N.Y. : New York Botanical Garden, 1935-; 1993 Jul. The
 Botanical review v. 59 (3): p. 155-210; 1993 Jul.  Includes
 references.
 
 Language:  English
 
 Descriptors: Aquatic plants; Allelopathy; Plant communities;
 Plant competition; Interactions; Literature reviews
 
 
 88                                   NAL Call. No.: 500 OK42
 Composition of essential oil from Proboscidea louisianica
 (Martyniaceae). Riffle, M.S.; Waller, G.R.; Murray, D.S.
 Oklahoma City, Okla. : The Academy; 1991.
 Proceedings of the Oklahoma Academy of Science v. 71: p.
 35-42; 1991. Includes references.
 
 Language:  English
 
 Descriptors: Oklahoma; Proboscidea louisianica;
 Allelochemicals; Essential oils; Plant composition
 
 
 89                                    NAL Call. No.: 4 AM34P
 Concentration dependency and stage of crop growth in alfalfa
 autotoxicity. Hegde, R.S.; Miller, D.A.
 Madison, Wis. : American Society of Agronomy; 1992 Nov.
 Journal of the American Society of Agronomy v. 84 (6): p.
 940-946; 1992 Nov. Includes references.
 
 Language:  English
 
 Descriptors: Medicago sativa; Crop residues; Allelochemicals;
 Shoots; Phytotoxicity; Seedlings; Seedling emergence; Phenolic
 compounds; Phytotoxins; Characterization; Identification
 
 Abstract:  Shoots of alfalfa (Medicago sativa L.) contain
 water-soluble chemical compounds which are autotoxic, i.e.,
 inhibit the growth of alfalfa itself. The objectives of this
 study were to (i) determine the inhibition threshold of the
 water-soluble chemical compounds (autotoxic principle), and
 (ii) demonstrate that the inhibition of seed germination on
 early seedling growth of alfalfa is due to autotoxinic
 compounds from alfalfa shoots and not from microbes. A farmer
 must make a decision on whether or not it would be advisable
 to replant alfalfa based on the level of alfalfa residue still
 growing in the field. Laboratory and greenhouse studies were
 conducted to determine if autotoxicity in alfalfa due to
 water-soluble compounds is concentration dependent. Filter-
 sterilized and non-filter-sterilized shoot aqueous extracts
 from vegetative and reproductive stages of 'WL-316' alfalfa
 were assayed at 20, 40, 60, and 80 g L-1 (fresh shoot weight
 basis) for their effect on seed germination and root and shoot
 elongation of seedlings of WL-316 alfalfa in a growth chamber.
 Compared with the control, root length, shoot length, and
 germination were inhibited beyond 20 g L-1 concentration.
 Shoot extract from the reproductive stage was more inhibitory
 than from the vegetative stage under laboratory conditions. In
 the greenhouse, incorporation of 4-wk-old green herbage from
 vegetative stage beyond 48 shoots per square meter level
 resulted in severe reductions in seedling emergence and plant
 fresh weight per unit area. Among the several phenolic
 compounds assayed for their phytotoxicity on root and shoot
 growth of alfalfa, coumarin and trans-cinnamic acid at 60 +/-
 10 micrograms mL-1 were the most inhibitory. Mixtures of five
 or more phenolic acids were more phytotoxic than their
 respective individual components except in the case of trans-
 cinnamic acid and coumarin. Autotoxicity in alfalfa may be
 caused by an interaction of many, yet uncharacterized chemical
 compounds present in shoots
 
 
 90                                  NAL Call. No.: 421 EN895
 The contribution of symbiotic yeast to toxin resistance of the
 cigarette beetle (Lasioderma serricorne).
 Dowd, P.F.; Shen, S.K.
 Dordrecht : Kluwer Academic Publishers; 1990 Sep.
 Entomologia experimentalis et applicata v. 56 (3): p. 241-248.
 ill; 1990 Sep. Includes references.
 
 Language:  English
 
 Descriptors: Lasioderma serRicorne; Larvae; Mortality;
 Detoxification; Allelochemicals; Flavonoids; Phenolic
 compounds; Phytotoxins; Resistance; Symbionts; Yeasts
 
 
 91                                  NAL Call. No.: SB610.W39
 Crop residue reduces jointed goatgrass (Aegilops cylindrica)
 seedling growth. Anderson, R.L.
 Champaign, Ill. : The Weed Science Society of America; 1993
 Jul. Weed technology : a journal of the Weed Science Society
 of America v. 7 (3): p. 717-722; 1993 Jul.  Includes
 references.
 
 Language:  English
 
 Descriptors: Colorado; Cabt; Triticum aestivum; Cultural weed
 control; Aegilops cylindrica; Allelopathy; Crop residues; Zea
 mays; Carthamus tinctorius; Panicum miliaceum; Sorghum
 bicolor; Nitrogen fertilizers; Immobilization; Nitrogen;
 Integrated control; Chemical control; Seedling stage;
 Triazinoneherbicides
 
 
 92                                    NAL Call. No.: QK1.C83
 Crop rotation.
 Bullock, D.G.
 Boca Raton, Fla. : CRC Press; 1992.
 Critical reviews in plant sciences v. 11 (4): p. 309-326;
 1992.  Literature review.  Includes references.
 
 Language:  English
 
 Descriptors: Rotations; Soil fertility; Cover crops;
 Sustainability; Soil organic matter; Soil structure; Erosion;
 Soil flora; Soil fauna; Insect pests; Allelopathy; Literature
 reviews
 
 
 93                                  NAL Call. No.: QH540.E23
 Crop rotation and intercropping strategies for weed
 management. Liebman, M.; Dyck, E.
 Tempe, Ariz. : Ecological Society of America; 1993 Feb.
 Ecological applications v. 3 (1): p. 92-122; 1993 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Weeds; Cultural weed control; Weed biology;
 Rotations; Intercropping; Seed banks; Plant density; Crop weed
 competition; Allelopathy
 
 
 94                                 NAL Call. No.: QD415.A1J6
 Defensive role of Allium sulfur compounds for leek moth
 Acrolepiopsis assectella Z. (Lepidoptera) against generalist
 predators. Nowbahari, B.; Thibout, E.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Nov.
 Journal of chemical ecology v. 18 (11): p. 1991-2002; 1992
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Allium; Plant composition; Defense;
 Allelochemicals; Acrolepiopsis assectella; Organic sulfur
 compounds; Volatile compounds; Formica fusca; Formica; Insect
 control
 
 Abstract:  It has been shown previously that sulfur volatiles
 produced by Allium plants affect the behavior of their
 specialist phytophages and of their specialist entomophages.
 The action of these compounds in protecting the leek moth
 Acrolepiopsis assectella against generalist entomophages was
 studied in comparison to the proposed original defensive role
 of these compounds against generalist herbivorous insects. Two
 ants species, Formica selysi and F. fusca, were used as
 generalist predators. Six behavioral criteria of the predatory
 behavior of the ants were studied in presence of the last-
 instar caterpillars (C5). C5 reared on artificial diets with
 or without leek components were tested, as well as C5 soaked
 in frass of leek-reared caterpillars or disulfide solutions.
 In addition, the response of the ants to pure chemicals found
 in leek was studied using honey solutions with or without
 sulfur compounds. The sulfur allelochemicals of Allium plants
 have a negative action on predatory ants. Interestingly, the
 nonvolatile precursors of sulfur volatiles of Allium plants
 seem to have a protective role for their phytophagous insects
 against generalist entomophages.
 
 
 95                                    NAL Call. No.: 4 AM34P
 Delayed seeding of alfalfa avoids autotoxicity after plowing
 or glyphosate treatment of established stands.
 Tesar, M.B.
 Madison, Wis. : American Society of Agronomy, [1949-; 1993
 Mar. Agronomy journal v. 85 (2): p. 256-263; 1993 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Michigan; Cabt; Medicago sativa; Allelopathy;
 Phytotoxicity; Continuous cropping; Sowing; Plowing;
 Glyphosate; Poa pratensis; Zea mays; Fallow; Seedlings;
 Density; Crop yield
 
 Abstract:  Re-establishment of alfalfa (Medicago sativa L.)
 has often been unsuccessful because of autotoxic effects of
 the crop on seedlings. This study was conducted at East
 Lansing, MI on an Udallic Ochraqualfs, fine loamy, mixed,
 mesic soil. The first objective was to determine the days
 required to eliminate autotoxic effects for l-, 4-, and 6-yr-
 old stands of alfalfa with 1.4 Mg ha-1 dry weight of topgrowth
 after treatments of plowing or glyphosate [N-
 (phosphonomethyl)glycine]. The second objective was to
 determine autotoxic effects of 1.4, 2.7, and 4.0 Mg ha-1 dry
 weight of topgrowth from year-old alfalfa plowed prior to
 seeding the same day. Check treatments were seedings after
 plowed Kentucky bluegrass (Poa pratensis L.), corn (Zea mays
 L.), or fallow. Alfalfa was seeded for Objective 1 on 23 June
 1982, 12 d after plowing alfalfa and 19 d after glyphosate
 application on 6-yr-old alfalfa; on 5 June 1984 and 29 May
 1985, 0, 7, 14, 21, 28, and 195 d (29 May 1985 only) after
 plowing alfalfa or glyphosate application on year-old alfalfa;
 and on 11 June 1986, 7, 14, 21, and 28 d after plowing alfalfa
 or glyphosate application on 4-yr-old alfalfa. Seedings for
 Objective 2 were made on 5 June 1984. Densities and yields of
 seedings equaled checks when made at least 14 d after plowing
 alfalfa or 21 d after glyphosate application on established
 alfalfa. Seedling densities were similar, but lower than the
 check, after three levels of alfalfa topgrowth were plowed and
 seeded the same day. Results indicate alfalfa can be re-
 established without significant autotoxicity if seedings are
 made at least 2 wk after plowing or 3 wk after glyphosate
 application on established alfalfa or after seeding failure.
 For maximum killing of old alfalfa and to avoid autotoxicity,
 plowing of alfalfa and seeding at least 2 wk after plowing, or
 early-fall or spring glyphosate application on alfalfa
 followed by no-till seeding at least 3 wk after glyphosate
 application, are recommended.
 
 
 96                                  NAL Call. No.: QK861.M63
 The determination of the allelopathic potential of pollen and
 nectar. Murphy, S.D.
 Berlin, W. Ger. : Springer-Verlag; 1992.
 Modern methods of plant analysis v. 13: p. 333-357; 1992.  In
 the series analytic: Plant toxin analysis / edited by H.F.
 Linskens and J.F. Jackson. Literature review.  Includes
 references.
 
 Language:  English
 
 Descriptors: Plant interaction; Allelopathins; Pollen; Nectar;
 Phytotoxicity; Isolation techniques; Bioassays; Chemical
 analysis; Literature reviews
 
 
 97                                  NAL Call. No.: 421 EN895
 Detoxification spectrum of the cigarette beetle symbiont
 Symbiotaphrina kochii in culture.
 Shen, S.K.; Dowd, P.F.
 Dordrecht : Kluwer Academic Publishers; 1991 Jul.
 Entomologia experimentalis et applicata v. 60 (1): p. 51-59;
 1991 Jul. Includes references.
 
 Language:  English
 
 Descriptors: Lasioderma serRicorne; Pesticide resistance;
 Symbionts; Yeasts; Hydrolases; Transferases; Allelochemicals;
 Microbial degradation; Mycotoxins; Pesticides; Detoxification
 
 
 98                                 NAL Call. No.: QD415.A1J6
 Devil's-claw (Proboscidea louisianica), essential oil and its
 components: potential allelochemical agents on cotton and
 wheat.
 Riffle, M.S.; Waller, G.R.; Murray, D.S.; Sgaramello, R.P. New
 York, N.Y. : Plenum Press; 1990 Jun.
 Journal of chemical ecology v. 16 (6): p. 1927-1940; 1990 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Proboscidea (martyniaceae); Essential oils;
 Chemical composition; Allelopathy; Germination; Phytotoxicity;
 Gossypium hirsutum; Triticumaestivum; Insect control;
 Biological control
 
 Abstract:  The potential allelopathic activity of devil's-claw
 [Proboscidea louisianica (Mill.) Thellung] essential oil and a
 few of the compounds it contains on the elongation of cotton
 (Gossypium hirsutum L.) and wheat (Triticum aestivum L.)
 radicles was studied using a Petri dish bioassay. Essential
 oil was collected by steam distillation using an all-glass-
 Teflon assembly. Ether extracts of the steam distillates from
 fresh devil's-claw were inhibitory to cotton and wheat radicle
 elongation. The following six components of devil's-claw
 essential oil identified by CGC-MS-DS were inhibitory to
 cotton and/or wheat at a concentration of 1 mM: vanillin,
 piperitenone, delta-cadinene, p-cymen-9-ol, alpha-bisabolol,
 and phenethyl alcohol.
 
 
 99                                 NAL Call. No.: QD415.A1J6
 Difference in hydroxamic acid content in roots and root
 exudates of wheat (Triticum aestivum L.) and rye (Secale
 cereale L.): possible role in allelopathy.
 Perez, F.J.; Ormeno-Nunez, J.
 New York, N.Y. : Plenum Press; 1991 Jun.
 Journal of chemical ecology v. 17 (6): p. 1037-1043; 1991 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Triticum aestivum; Secale cereale; Root exudates;
 Allelopathy; Chemical composition; Bioassays; Weed control;
 Biological control
 
 Abstract:  Hydroxamic acids (Hx) produced by some cereal crops
 have been associated with allelopathy. However, the release of
 Hx to the soil by the producing plant--an essential condition
 for a compound to be involved in allelopathy--has not been
 shown. GC and HPLC analysis of roots and root exudates of
 wheat (Triticum aestivum L.) and rye (Secale cereale L.)
 cultivars, with high Hx levels in their leaves, demonstrated
 the presence of these compounds in the roots of all cultivars
 analyzed and in root exudates of rye. Moreover, bioassays
 employing root exudates collected from wheat and rye seedlings
 demonstrated that only rye exudates inhibited root growth of
 wild oats, Avena fatua L., a weed whose root growth is
 inhibited by Hx. These results suggest that rye could
 potentially interfere with the growth of Avena fatua in nature
 and that this interference could be due to the release of Hx
 to the soil by way of roots.
 
 
 100                                  NAL Call. No.: QP501.C6
 Differences in cytochrome p450 activities in tobacco budworm
 larvae as influenced by resistance to host plant
 allelochemicals and induction. Rose, R.L.; Gould, F.; Levi,
 P.E.; Hodgson, E.
 Oxford : Pergamon Press; 1991.
 Comparative biochemistry and physiology : B : Comparative
 biochemistry v. 99 (3): p. 535-540; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Heliothis virescens; Larvae; Strains; Cytochrome
 p-450; Oxygenases; Isoenzymes; Enzyme activity; 2-tridecanone;
 Nicotine; Quercetin; Resistance; Induction; Resistance
 mechanisms; Oxidation; Metabolism; Substrates
 
 Abstract:  1. Nicotine and 2-tridecanone resistant strains of
 the tobacco budworm, Heliothis virescens (F), had elevated
 cytochrome P450 content and significant increases in
 metabolism of five of six monooxygenase substrates relative to
 two susceptible strains. 2. Resistance to quercetin did not
 result in an increase in cytochrome P450 content; however,
 significant increases in metabolism were observed for two
 monooxygenase substrates. 3. P450 content was significantly
 induced by nicotine and 2-tridecanone, but not by quercetin.
 4. Patterns of substrate oxidations varied between strains and
 inducing agents, suggesting that different isozymes of P450
 are associated with resistance and induction.
 
 
 101                                NAL Call. No.: QD415.A1J6
 Differential allelochemical detoxification mechanism in tissue
 cultures of Antennaria microphylla and Euphorbia esula.
 Hogan, M.E.; Manner, G.D.
 New York, N.Y. : Plenum Press; 1991 Jan.
 Journal of chemical ecology v. 17 (1): p. 167-174; 1991 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Antennaria microphylla; Euphorbia esula; Cell
 suspensions; Callus; Allelochemicals; Hydroquinone; Metabolic
 detoxification
 
 Abstract:  Callus and suspension cultures of Antennaria
 microphylla (small everlasting) and the noxious weed Euphorbia
 esula (leafy spurge) can glucosylate benzene-1,4-diol
 (hydroquinone) to the corresponding monoglucoside, arbutin.
 HPLC analysis of extracts from callus tissue corroborates the
 presence of hydroquinone in the cells of small everlasting.
 Constitutive levels of a UDPG-dependent glucosyltransferase
 were detected in cell-free extracts of this tissue. Although
 this detoxification enzyme was induced in leafy spurge
 suspension culture cells grown in the presence of
 hydroquinone, the activity was six-fold lower than that
 measured in small everlasting. Differential ability to
 detoxify hydroquinone provides a basis for the observed
 allelopathic interaction between small everlasting and leafy
 spurge.
 
 
 102                                  NAL Call. No.: 79.8 W41
 Differential inhibition of seed germination by sweetpotato
 (Ipomoea batatas) root periderm extracts.
 Peterson, J.K.; Harrison, H.F. Jr
 Champaign, Ill. : Weed Science Society of America; 1991 Jan.
 Weed science v. 39 (1): p. 119-123; 1991 Jan.  Includes
 references.
 
 Language:  English
 
 Descriptors: Ipomoea batatas; Competitive ability; Abutilon
 theophrasti; Amaranthus retroflexus; Cassia occidentalis;
 Eclipta alba; Eleusine indica; Pharbitis purpurea; Panicum
 miliaceum; Solanum nigrum; Seed germination; Germination
 inhibitors; Allelopathins; Periderm; Sweet potato extract;
 Bioassays; Allelopathy; Crop weed competition
 
 Abstract:  The effect of sequential hexane, ethyl acetate, and
 aqueous methanol extracts of 'Regal' sweetpotato periderm on
 seed germination of sweetpotato, proso millet, and seven weed
 species was studied. The hexane extract, which contained the
 nonpolar components of the periderm tissue, was least
 inhibitory. It inhibited velvetleaf, proso millet, black
 nightshade, and redroot pigweed germination, and maximum
 inhibition was 56% for black nightshade at 200 mg of periderm
 extracted ml-1. The ethyl acetate fraction was inhibitory to
 proso millet, velvetleaf, black nightshade, goosegrass, tall
 morningglory, coffee senna, and redroot pigweed. The estimated
 I50(3) for ethyl acetate ranged from 17 mg periderm extracted
 ml-1 for black nightshade to 201 mg ml-1 for coffee senna.
 Sweetpotato, tall morningglory, and eclipta germination was
 not inhibited by this extract at the concentrations tested.
 The aqueous methanol extract was much more inhibitory than the
 hexane or ethyl acetate extracts, and there was considerable
 variation between species in response to this extract The I50
 estimates for the aqueous methanol extract were 0.5, 0.6, 2.8,
 4.4, 5.1, 9.6, 15.7, 21.0, and 25.8 mg ml-1 for velvetleaf,
 proso millet, black nightshade, goosegrass, sweetpotato, tall
 morningglory, eclipta, coffee senna, and pigweed,
 respectively.
 
 
 103                                  NAL Call. No.: 23 AU783
 Differential response of wheat to retained crop stubbles. I.
 Effect of stubble type and degree of decomposition.
 Purvis, C.E.
 Melbourne : Commonwealth Scientific and Industrial Research
 Organization; 1990.
 Australian journal of agricultural research v. 41 (2): p.
 225-242. ill; 1990. Includes references.
 
 Language:  English
 
 Descriptors: New South Wales; Triticum aestivum; Sowing date;
 Yield response functions; Phytotoxicity; Stubble mulching;
 Allelopathy
 
 
 104                                  NAL Call. No.: 23 AU783
 Differential response of wheat to retained crop stubbles. II.
 Other factors influencing allelopathic potential;
 intraspecific variation, soil type and stubble quantity.
 Purvis, C.E.; Jones, G.P.D.
 Melbourne : Commonwealth Scientific and Industrial Research
 Organization; 1990.
 Australian journal of agricultural research v. 41 (2): p.
 243-251; 1990. Includes references.
 
 Language:  English
 
 Descriptors: New South Wales; Triticum aestivum; Yield
 response functions; Allelopathy; Growth; Inhibition;
 Phytotoxicity; Seedling emergence; Soil types; Stubble
 mulching
 
 
 105                                NAL Call. No.: QD415.A1J6
 Do defoliation and subsequent phytochemical responses reduce
 future herbivory on oak trees?.
 Faeth, S.H.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Jun.
 Journal of chemical ecology v. 18 (6): p. 915-925; 1992 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Quercus; Defoliation; Responses; Chemical
 composition; Leaves; Tannins; Protein content;
 Allelochemicals; Defense; Chemical ecology
 
 Abstract:  Perennial plants are thought to respond to partial
 or complete defoliation by producing new foliage that is less
 susceptible to herbivores because of induction of
 allelochemicals. Here, I tested this hypothesis by manually
 removing primary foliage from branches of Quercus emoryi
 (Fagaceae) at two different times in the season and monitoring
 changes in protein and tannin levels and the amount of
 herbivory relative to control branches. New, secondary leaves
 had 2.5 X greater hydrolyzable tannin content than mature
 foliage of control branches. Condensed tannins, which
 constitute a relatively low fraction of leaf mass, were lower,
 while protein content was temporarily greater, in new
 secondary leaves relative to mature leaves. Despite large
 increases in hydrolyzable tannins, herbivory levels were
 greater on refoliated branches than on control branches. New
 foliage is susceptible to herbivory regardless of when it is
 produced in the season, possibly because lower toughness and
 higher water content override any induced or developmentally
 related changes in allelochemistry. My results do not support
 the hypothesis that postherbivore changes in phytochemistry
 protect perennial plants from future herbivory, at least
 within a growing season.
 
 
 106                                  NAL Call. No.: 450 J829
 Dynamics of associations between plants in ten old fields
 during 31 years of succession.
 Myster, R.W.; Pickett, S.T.A.
 Oxford : Blackwell Scientific; 1992.
 Journal of ecology v. 80 (2): p. 291-302; 1992.  Includes
 references.
 
 Language:  English
 
 Descriptors: New Jersey; Plant succession; Old fields;
 Community ecology; Plant ecology; Plant competition;
 Allelopathy
 
 Abstract:  The pattern of significant associations between
 plants was examined in ten old fields during 31 years of
 succession by calculating rank correlations for species pairs
 in each old field during each sample year. Three hypotheses
 were tested concerning the dynamics of species interactions
 through succession, and correspondence was explored between
 the pattern of association and published results from field
 and glasshouse experiments. The proportion, number and level
 of significance of associations between plants all declined
 with time. Annuals and biennials had a higher portion of
 significant associations and more positive associations than
 perennial species. Plant species involved in many, significant
 associations and implicated as actively interacting with other
 species were generally neither native nor the most abundant.
 Seventy per cent of the species analysed in the present study,
 that had also been used in field and glasshouse experiments
 demonstrating competition reported in the literature, were
 involved in significant and repeated negative pairwise
 associations. However, only 33% of species used in field and
 glasshouse experiments demonstrating allelopathy showed such
 correspondence. Grasses may be major inhibitory species
 because they were involved in many significant negative plant
 associations although they did not achieve high abundance in
 these old fields. Lonicera japonica and Rosa multiflora were
 woody species involved in many negative associations and may
 play major roles by inhibiting later successional species.
 Investigations into the role of species interactions during
 succession may focus productively on those relatively few
 species that are strongly associated. The timing and the
 consequences of these associations may illumunate how
 interaction mechanisms such as competition and alleopathy
 structure successions.
 
 
 107                                  NAL Call. No.: 450 AM36
 The ecological impact of allelopathy in Ailanthus altissima
 (Simaroubaceae). Lawrence, J.G.; Colwell, A.; Sexton, O.J.
 Columbus, Ohio : Botanical Society of America; 1991 Jul.
 American journal of botany v. 78 (7): p. 948-958; 1991 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Missouri; Ailanthus altissima; Allelopathy; Plant
 communities; Growth inhibitors; Species diversity; Toxicity;
 Environmental factors
 
 Abstract:  Compounds inhibitory to the growth of neighboring
 plant species were found in significant concentrations in the
 leaves and stems of young Ailanthus altissima ramets. The
 surrounding soil also contained appreciable concentrations of
 similarly acting toxins. Individuals of neighboring plant
 species have either incorporated active portions of inhibitory
 compounds or responded to Ailanthus by producing growth-
 inhibiting substances. Under greenhouse conditions,
 individuals of neighboring plant species previously unexposed
 to Ailanthus in the field were found to be more susceptible to
 the Ailanthus toxins than individuals previously exposed.
 Moreover, seeds produced by unexposed populations were also
 more susceptible to Ailanthus toxins than seeds produced by
 previously exposed populations. These differences demonstrated
 that the allelochemicals of Ailanthus altissima exhibited a
 measurable impact upon neighboring plant species. Since the
 progeny of these populations displayed a differential response
 to Ailanthus toxin, this phenotypic difference between the two
 populations may have a heritable basis.
 
 
 108                                NAL Call. No.: QD415.A1J6
 Effect of apiforol and apigeninidin on growth of selected
 fungi. Schutt, C.; Netzly, D.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2261-2266; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Sorghum bicolor; Disease resistance; Allelopathy;
 Seeds; Plant composition; Antifungal properties
 
 Abstract:  Selected fungi were grown on agar plates in the
 presence of naringenin, apiforol, apiforol 7-O-
 rhamnoglucoside, or apigeninidin. Of the four compounds
 tested, only apigeninidin inhibited the growth of Fusarium
 oxysporum, Gibberella zeae, Gliocladium roseum, Alternaria
 solani, and Phytophthora infestans. In contrast, the growth of
 Rhizoctonia solani, Sclerotium rolfsii, and Rhizopus
 stolonifer (- and +) was not effected by any compound. Since
 apigeninidin is present in seeds of Sorghum sp., we
 hypothesize that apigeninidin may play a role in mold
 resistance and that apiforol accumulates as a biosynthetic
 precursor of apigeninidin, not as a fungal defense compound.
 
 
 109                                NAL Call. No.: QD415.A1J6
 Effect of diacetyl piquerol on H+-ATPase activity of
 microsomes from Ipomoea purpurea.
 Cruz Ortega, R.; Anaya, A.L.; Gavilanes-Ruiz, M.; Sanchez
 Nieto, S.; Jimenez Estrada, M.
 New York, N.Y. : Plenum Press; 1990 Jul.
 Journal of chemical ecology v. 16 (7): p. 2253-2261; 1990 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Pharbitis purpurea; Leaves; Microsomes;
 Adenosinetriphosphatase; Enzyme activity; Proton pump;
 Allelopathy; Bioassays; Inhibition; Weedcontrol
 
 Abstract:  The effect of an allelopathic compound, diacetyl-
 piquerol on the H+ -ATPase activity of the microsomal fraction
 from the radicles of a common weed lpomoea purpurea was
 studied. The diacetyl-piquerol inhibited the germination and
 radicle growth from I. purpurea; the radicle growth was
 increasingly inhibited (10% to 100%) as piquerol
 concentrations were raised (10 micromole to 1000 micromole).
 The H+ -ATPase activity was inhibited (48%) by 500 micromole
 diacetyl-piquerol, and this inhibition was higher in plasma
 membrane ATPase (67.2%) than in tonoplast membrane ATPase
 (31.4%). Additional studies of the precise physiological
 mechanisms of interference caused by allelopathic compounds
 are needed.
 
 
 110                                  NAL Call. No.: SD13.C35
 The effect of host variability on growth and performance of
 the introduced pine sawfly, Diprion similis.
 Codella, S.G. Jr; Fogal, W.H.; Raffa, K.F.
 Ottawa, Ont. : National Research Council of Canada; 1991 Nov.
 Canadian journal of forest research; Journal canadien de
 recherche forestiere v. 21 (11): p. 1668-1674; 1991 Nov. 
 Includes references.
 
 Language:  English
 
 Descriptors: Pinus banksiana; Pinus strobus; Diprion similis;
 Larvae; Host range; Growth; Foliage; Leaf age; Nutrient
 content; Allelochemicals; Fecundity; Feeding behavior
 
 Abstract:  Diprion similis (Htg.) (Hymenoptera: Diprionidae)
 was reared in the laboratory on Pinus banksiana Lamb. and
 Pinus strobus L. from the second stadium through adult
 emergence. Groups of larvae were fed current-year or previous
 years' foliage from specific trees. Host species had a
 significant, but limited, effect on D. similis growth and
 performance. Foliar age had a stronger and more consistent
 effect on development. In contrast with reports for other
 diprionid species, previous years' needles had a consistently
 greater adverse effect on D. similis performance than did new
 growth. This suggests that there is a conflict between the
 avoidance of host tissues with high allelochemical
 concentrations and the avoidance of those with reduced
 nutrient content. Larval survival did not vary between foliar
 treatments, which suggests that the detrimental effects of
 host diet are chronic rather than acute. Female fecundity was
 strongly associated with cocoon weight, but the relationship
 varied with host diet and diapause incidence, as did the
 relative reproductive potential. Substantial between-tree
 variability in insect performance indicates a potential for
 resistance breeding programs. Tree rankings for each
 performance variable were highly consistent, which would
 permit the development of an expeditious screening procedure.
 
 
 111                                   NAL Call. No.: S295.M3
 Effect of maize residue on five maize hybrids.
 Wahab, Z.B.; Kaspar, T.C.
 Selangar : Malaysian Society of Applied Biology; 1990 Jun.
 Malaysian applied biology : Biologi gunaan Malaysia v. 19 (1):
 p. 29-36; 1990 Jun.  Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Hybrids; Rotations; Varietal reactions;
 Varietal tolerance; Crop residues; Maize; Allelopathy;
 Allelochemicals; Seedlings; Growth; Inhibition; Leaves;
 Height; Roots; Shoots; Dry matter; Weight; Ratios; Crop growth
 stage; Crop yield; Crop losses
 
 
 112                                  NAL Call. No.: 470 AM36
 The effect of nutrients and enriched CO2 environments on
 production of carbon-based allelochemicals in Plantago: a test
 of the carbon/nutrient balance hypothesis.
 Fajer, E.D.; Bowers, M.D.; Bazzaz, F.A.
 Chicago, Ill. : University of Chicago Press; 1992 Oct.
 The American naturalist v. 140 (4): p. 707-723; 1992 Oct. 
 Includes references.
 
 Language:  English
 
 Descriptors: Plantago lanceolata; Carbon dioxide; Secondary
 metabolites; Aucubin; Genotypes; Plant nutrition
 
 
 113                                   NAL Call. No.: QK1.A28
 Effect of staling product from selected phylloplane fungi on
 in vitro growth of two pathogenic fungi.
 Singh, D.B.
 Meerut, India : Society for Advancement of Botany; 1990 Dec.
 Acta botanica Indica v. 18 (2): p. 256-259; 1990 Dec. 
 Includes references.
 
 Language:  English
 
 Descriptors: Brassica nigra; Hordeum vulgare; Phylloplane
 fungi; Alternaria brassicae; Drechslera; Culture filtrates;
 Plant disease control; Allelopathins; Metabolites; Growth
 inhibitors
 
 
 114                                NAL Call. No.: QD415.A1J6
 Effect of temperature and sucrose concentration on
 hydroquinone toxicity in leafy spurge suspension culture cells.
 Hogan, M.E.; Manners, G.D.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Sep.
 Journal of chemical ecology v. 18 (9): p. 1541-1549; 1992 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Euphorbia esula; Callus; Cell suspensions;
 Hydroquinone; Phytotoxicity; Sucrose; Temperature; Metabolic
 detoxification; Allelopathy; Weed control; Antennaria
 microphylla
 
 Abstract:  Euphorbia esula (leafy spurge) suspension culture
 cell bioassays were used to determine whether sucrose
 accumulation enhanced the glucosylation (detoxification) of
 hydroquinone in this noxious weed. The bioassay results
 indicate that cold temperatures and exogenous hydroquinone
 represent a dual stress to spurge cell growth that can be
 partially ameliorated by hydrolysis of sucrose. The persistent
 susceptibility of leafy spurge suggests that hydroquinone-
 producing forage plants (which are not toxic to animals) might
 be used as natural competitors.
 
 
 115                                NAL Call. No.: QD415.A1J6
 Effects of cotton plant allelochemicals and nutrients on
 behavior and development of tobacco budworm.
 Hedin, P.A.; Parrott, W.L.; Jenkins, J.N.
 New York, N.Y. : Plenum Press; 1991 Jun.
 Journal of chemical ecology v. 17 (6): p. 1107-1121; 1991 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Gossypium hirsutum; Pest resistance; Heliothis
 virescens; Allelochemicals; Terpenoids; Gossypol;
 Interactions; Amino acids; Insect control; Biological control
 
 Abstract:  Female moths of the tobacco budworm, Heliothis
 virescens (F.), oviposit in the terminals of the cotton plant,
 Gossypium hirsutum (L.). The hatched larvae migrate to the
 terminal area and then to small squares (buds), on which they
 feed, finally burrowing into the anthers where they grow and
 develop. They attempt to avoid gossypol glands as they feed.
 Chemically related evidence explains, in part, these
 observations. The calyx crown of resistant lines (which is
 avoided) is high in the terpenoid aldehydes (TAs) including
 gossypol. HPLC data showed that the gossypol content of both
 susceptible and resistant glanded lines is equal, while the
 hemigossypolone and heliocides H1 and H2 are greatly increased
 in resistant lines and presumably are more closely associated
 with resistance. Analysis for total amino acids in cotton
 square tissues showed that there was a gradation from the
 calyx and calyx crown, which were lowest, to the anthers, the
 site of final insect development, which were highest.
 Synthetic diets mimicking amino acid distribution in anthers
 were found to be successful for larval growth and development.
 
 
 116                                NAL Call. No.: QD415.A1J6
 Effects of exogenously applied ferulic acid, a potential
 allelopathic compound, on leaf growth, water utilization, and
 endogenous abscisic acid levels of tomato, cucumber, and bean.
 Holappa, L.D.; Blum, U.
 New York, N.Y. : Plenum Press; 1991 May.
 Journal of chemical ecology v. 17 (5): p. 865-886; 1991 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lycopersicon esculentum; Cucumis sativus;
 Phaseolus vulgaris; Allelopathy; Ferulic acid; Abscisic acid
 
 Abstract:  To determine the relative sensitivities of tomato,
 cucumber, and bean to exogenously applied concentrations of
 ferulic acid (FA) and to determine whether FA-induced stress
 responses increase endogenous levels of abscisic acid (ABA),
 wild-type and flacca (ABA-deficient mutant) tomato
 (Lycopersicon esculentum Mill. cv. Ailsa Craig), cucumber,
 (Cucumis sativus L. cv. Early Green Cluster), and bean
 (Phaseolus vulgaris L. cv. Oregon 91) were treated with FA
 (0.0, 0.2, 0.4, 0.8 mM) in nutrient solution every other day
 for a total of two or three treatments. FA inhibited leaf
 growth and water utilization of wild-type tomato, flacca
 tomato, and cucumber, but not of bean. Acclimation to FA was
 observed following the first FA treatment and increased
 endogenous ABA levels were found in wild-type tomato, flacca
 tomato, and cucumber following multiple FA treatments.
 Induction of ABA biosynthesis occurred in wild-type tomato
 within 8 hr of FA treatment and maximum ABA levels were
 observed 24 hr after treatment. At that time, ABA levels of
 tomato treated with 0.4 and 0.8 mM FA were 13.7 times and 2.6
 times higher than control levels, respectively. A second FA
 (0.4 or 0.8 mM) treatment, 48 hr after the first, did not
 appear to affect ABA levels. Ninety-six hours after the first
 treatment, ABA levels of tomato treated with 0.4 mM FA
 approached control levels; ABA levels of plants treated with
 0.8 mM FA were 1.9 times higher than control levels. Control
 ABA levels increased gradually with time. The data showed that
 plant sensitivity and ability of subsequent acclimation to
 phenolic acids, such as FA, were taxa dependent.
 
 
 117                                NAL Call. No.: QD415.A1J6
 Effects of ferulic acid, and allelopathic compound, on net P,
 K, and water uptake by cucumber seedlings in a split-root
 system.
 Lyu, S.W.; Blum, U.
 New York, N.Y. : Plenum Press; 1990 Aug.
 Journal of chemical ecology v. 16 (8): p. 2429-2439; 1990 Aug. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Seedlings; Water uptake; Ferulic
 acid; Allelopathy
 
 Abstract:  Since distribution of allelopathic compounds in
 soils is highly variable, injurious effects by such compounds
 should be related to the frequency of contact with roots.
 Experiments were conducted to determine how P, K, and water
 uptake of cucumber seedlings were affected as the fraction of
 roots in contact with ferulic acid (FA) was increased.
 Seedlings were grown in Hoagland's nutrient solution for 14
 days and then transferred to 0.5 mM CaSO4 solution for 24 hr
 before being placed into a split-root culture system. The
 containers in the system were filled with 0.5 mM
 concentrations of KH2PO4 and CaSO4 or 0.5 mM concentrations of
 KH2PO4, CaSO4, and ferulic acid (FA). Net uptake of P by
 seedlings (milligrams per seedling) decreased in a curvilinear
 (concave) manner as the fraction of the roots in contact with
 FA increased. Net uptake of K (milligrams per seedling) and
 water (milliliters per seedling) by seedlings decreased
 linearly as the fraction of the roots in contact with FA
 increased. Net uptake of P, K, and water by seedlings was
 reduced 57, 75, and 29%, respectively, when the whole root
 system was exposed to FA. Net P and K uptake of roots
 (milligrams per gram root fresh weight) not in contact with FA
 decreased in a linear and curvilinear (convex) manner,
 respectively, as the fraction of roots in contact with FA
 increased. Net P and K uptake of roots in contact with ferulic
 acid increased in a linear and curvilinear (convex) manner,
 respectively. Net water uptake of roots (milliliters per gram
 root fresh weight) not in contact with FA increased in a
 curvilinear (concave) manner as the frequency of the roots in
 contact with FA increased. Net water uptake of roots in
 contact with FA did not show a trend. Transpiration
 (milliliters per square centimeter) was reduced in a linear
 manner as the fraction of roots in contact with FA increased.
 A very slight compensation by roots not in contact with FA for
 roots in contact with FA was observed for net water uptake
 rate
 
 
 118                                NAL Call. No.: QD415.A1J6
 Effects of juglone on growth, photosynthesis, and respiration.
 Hejl, A.M.; Einhellig, F.A.; Rasmussen, J.A.
 New York, N.Y. : Plenum Publishing Corporation; 1993 Mar.
 Journal of chemical ecology v. 19 (3): p. 559-568; 1993 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lemna minor; Glycine max; Juglone; Allelopathy;
 Photosynthesis; Chloroplasts; Mitochondria; Respiration
 
 Abstract:  The impacts of juglone on plant growth and several
 other physiological functions were evaluated in this study.
 Juglone inhibited Lemna minor growth, chlorophyll content, and
 net photosynthesis at treatments between 10 and 40 micromolar.
 Soybean leaf disks vacuum infiltrated with as little as 10
 micromolar juglone had reduced photosynthesis. Oxygen
 evolution by chloroplasts isolated from Pisum sativum was
 inhibited by juglone with an I50 of 2 micromolar. Micromolar
 treatments of juglone stimulated oxygen uptake in mitochondria
 isolated from Glycine max. These data suggest perturbations of
 chloroplast and mitochondrial functions may contribute to
 plant growth reductions observed in juglone-mediated
 allelopathy.
 
 
 119                                 NAL Call. No.: 381 J8223
 Effects of kinetin formulations on allelochemicals and
 agronomic traits of cotton.
 Hedin, P.A.; McCarty, J.C. Jr
 Washington, D.C. : American Chemical Society; 1991 Mar.
 Journal of agricultural and food chemistry v. 39 (3): p.
 549-553; 1991 Mar. Includes references.
 
 Language:  English
 
 Descriptors: Gossypium hirsutum; Allelochemicals; Plant growth
 regulators; Plantcomposition; Plant analysis; Yield increases
 
 Abstract:  Twelve candidate plant growth regulator
 formulations were applied twice at two levels to fruiting
 cotton (Gossypium hirsutum L.). Leaves and squares were
 collected for analysis of allelochemicals (gossypol, tannin,
 anthocyanin, flavonoids) at 3 and 5 weeks after the first
 treatment. The plots were machine harvested one time to
 determine yield. Seeds were delinted and analyzed for
 agronomic traits and gossypol. Leaf gossypol and square
 gossypol were the categories most frequently increased by the
 bioregulators. Kinetin, and kinetin plus CaCl2 or Na2SeO3, and
 mepiquat chloride (PIX) alone or with a commercial cytokinin
 preparation (Foliar Triggrr) all increased gossypol and one or
 more of the other allelochemicals significantly. A sugar-amino
 acid fraction isolated from Foliar Triggrr increased cotton
 yield by 26% when applied as a foliar dressing at 2.88 mol/ha,
 as well as increasing gossypol. These results suggest that
 plants under stress may respond positively to nutrient foliar
 applications, giving both increased allelochemicals and
 improved yield.
 
 
 120                                NAL Call. No.: QD415.A1J6
 Effects of mixtures of four phenolic acids on leaf area
 expansion of cucumber seedlings grown in Portsmouth B soil
 materials.
 Gerig, T.M.; Blum, U.
 New York, N.Y. : Plenum Press; 1991 Jan.
 Journal of chemical ecology v. 17 (1): p. 29-40; 1991 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Seedlings; Leaves; Phenolic
 acids; Allelopathy; Soil treatment; Leaf area
 
 Abstract:  Cucumber seedlings growing in a 1:2 mixture of soil
 (Portsmouth B1) and sand adjusted to pH 5.2 were treated every
 other day five times with 0, 0.0625, 0.125, 0.25, or 0.5
 micromoles/g soil of ferulic, caffeic, p-coumaric, p-
 hydroxybenzoic, protocatechuic, sinapic, syringic, or vanillic
 acids. Treatments began when seedlings were 8 days old. The
 effects on mean absolute rates of leaf expansion were used to
 estimate the relative potencies of these phenolic acids to
 ferulic acid. Based on the results of this experiment,
 ferulic, p-coumaric, p-hydroxybenzoic, and vanillic acids were
 chosen for further study. Materials and procedures were
 identical in the second study, but treatments consisted of
 mixtures of the four phenolic acids at concentration
 combinations designed to achieve 40% or 60% inhibition of
 absolute rates of leaf expansion. Using joint action analysis,
 a model describing the action of the phenolic acid mixtures
 was developed. A model involving only two factor terms was
 sufficient to describe the observed responses of cucumber leaf
 area to the phenolic acid mixtures. The action of p-
 hydroxybenzoic acid on absolute rates of leaf expansion was
 inhibited by the presence of the other three phenolic acids.
 No other antagonisms or synergisms existed among the four
 compounds.
 
 
 121                                   NAL Call. No.: QR1.F44
 Effects of phenolic acids on ammonia oxidation by terrestrial
 autrotrophic nitrifying microorganisms.
 McCarty, G.W.; Bremner, J.M.; Schmidt, E.L.
 Amsterdam : Elsevier Science Publishers; 1991 Jul.
 FEMS microbiology letters - Federation of European
 Microbiological Societies v. 85 (4): p. 345-350; 1991 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Nitrosomonas; Nitrosolobus; Nitrobacteraceae;
 Nitrification; Ammonia; Oxidation; Nitrites; Caffeic acid; P-
 coumaric acid; Ferulic acid; Vegetation; Allelopathy
 
 Abstract:  It has been hypothesized that vegetation in certain
 ecosystems inhibits nitrification in soil by producing
 phenolic compounds that inhibit oxidation of ammonia by
 nitrifying microorganisms. This hypothesis is based largely on
 a report that very low concentrations (10(-6) M-10(-8) M) of
 several phenolic acids (notably ferulic acid) completely
 inhibited NO2-production in an aqueous suspension of soil
 treated with (NH4)2SO4 and a nutrient solution suitable for
 growth of Nitrosomonas and other autotrophic nitrifying
 microorganisms. To evaluate this hypothesis, we determined the
 effects of three phenolic acids (ferulic acid, caffeic acid,
 and p-coumaric acid) on nitrite production by representatives
 of three genera of terrestrial autotrophic nitrifying
 microorganisms (Nitrosospira, Nitrosomonas, or Nitrosolobus)
 grown on a defined medium containing NH4+. We found that
 nitrite production by the Nitrosospira was not inhibited by
 ferulic acid, caffeic acid, or p-coumaric acid at
 concentrations of 10(-6) or 10(-5) M and was only slightly
 inhibited when these acids were at a concentration of 10(-4)
 M. We also found that ferulic acid did not markedly inhibit
 nitrite production by the three genera of nitrifying
 microorganisms studied, even when its concentration was as
 high as 10(-3) M. These observations invalidate the hypothesis
 tested because the phenolic acids studied did not
 significantly retard ammonia oxidation by autotrophic
 microorganisms even when their concentration in cultures of
 these microorganisms greatly exceeded their concentrations in
 soils.
 
 
 122                                 NAL Call. No.: QK745.P55
 Effects of phenolic acids on germination and seedling growth
 of corn (Zea mays), radish (Raphanus sativus), and peanut
 (Arachis hypogaea). El Abdaoui, F.; Foy, C.L.
 Frederick, Md.: Plant Growth Regulator Society of America;
 1993 Apr. Quarterly - PGRSA v. 21 (2): p. 49-63; 1993 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Raphanus sativus; Arachis hypogaea;
 Seed germination; Seedling growth; Phytotoxicity; Gallic acid;
 Vanillic acid; Ferulic acid; Allelochemicals
 
 
 123                               NAL Call. No.: QL495.A1I57
 Effects of plant flavonoids and other allelochemicals on
 insect cytochrome P-450 dependent steroid hydroxylase
 activity.
 Mitchell, M.J.; Keogh, D.P.; Crooks, J.R.; Smith, S.L.
 Exeter : Pergamon Press; 1993 Jan.
 Insect biochemistry and molecular biology v. 23 (1): p. 65-71;
 1993 Jan. Special issue: Ecdysone: from biosynthesis to
 regulation of gene expression. Papers from the Xth Ecdysone
 Workshop, April 6-9, 1992, Liverpool, England. Includes
 references.
 
 Language:  English
 
 Descriptors: Plant extracts; Allelochemicals; Flavonoids;
 Insecticidal action; Aedes aegypti; Drosophila melanogaster;
 Manduca sexta; Molting hormones; Cytochrome p-450; Enzyme
 activity; Oxygenases
 
 Abstract:  The plant flavonoids flavone, chrysin, apigenin,
 kaempferol, morin, quercetin, myricetin and phloretin were
 found to inhibit in a dose-dependent manner the cytochrome
 P450 dependent ecdysone 20-monooxygenase activity associated
 with adult female Aedes aegypti, wandering stage larvae of
 Drosophila melanogaster, and fat body and midgut from
 prewandering and wandering stage last instar larvae of Manduca
 sexta. The concentrations of these flavonoids required to
 elicit a 50% inhibition of the steroid hydroxylase activity in
 all the insects ranged from ca 1 X 10(-6) to 1 X 10(-5) M. In
 addition, lower concentrations (1 X 10(-6) to 1 X 10(-6) M) of
 the flavonoids kaempferol, morin, quercetin and myricetin
 significantly stimulated (50-100% above control) M. sexta fat
 body ecdysone 20-monooxygenase activity. Other plant
 allelochemicals examined and found to significantly inhibit
 insect ecdysone 20-monooxygenase activity include
 corynanthine, quinidine, and quinine; whereas, indican and
 mimosine were found to significantly stimulate M. sexta fat
 body steroid hydroxylase activity. Several allelochemicals
 were without effect at all concentrations tested. Although
 none of the compounds tested in this study elicited effects at
 very low concentrations (1 X 10(-9) to 1 X 10(-8) M), the in
 vitro monooxygenase radioassay does hold considerable promise
 as a screening toot for the detection and identification of
 plant allelochemicals which may function as biopesticides
 affecting insect ecdysteroidogenesis.
 
 
 124                                NAL Call. No.: QD415.A1J6
 Effects of salicylic acid on growth and stomatal movements of
 Vicia faba L.: evidence for salicylic acid metabolization.
 Manthe, B.; Schulz, M.; Schnabl, H.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Sep.
 Journal of chemical ecology v. 18 (9): p. 1525-1539; 1992 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Vicia faba; Salicylic acid; Allelopathy; Growth;
 Shoots; Sprout inhibition; Metabolic detoxification;
 Transpiration
 
 Abstract:  The influence of salicylic acid on the growth and
 stomatal movements of Vicia faba L. was investigated. Whereas
 shoot length, fresh weight, and transpiration rates, which are
 directly correlated with stomatal pore widths, were only
 affected at salicylic acid concentrations higher than 3.5 mM
 after long-term treatments, guard cells in epidermal peels
 exhibited a high sensitivity at concentrations as low as 0.001
 mM, resulting in stomatal closing. HPLC analysis of methanolic
 extracts from roots and leaves revealed the presence of free
 salicylic acid and a metabolite, whose amount increased with
 time in plants previously incubated with a medium containing
 salicylic acid. The possible ability of Vicia faba to detoxify
 the phenolic acid may be one explanation of the discrepancy
 between the stomatal reaction in epidermal peels directly
 treated with the phenolic acid and after application through
 the transpiration stream. The results indicate that, under
 natural conditions, salicylic acid will not act as an
 allelopathic compound whose toxic properties severely affect
 the growth of Vicia faba.
 
 
 125                                 NAL Call. No.: 60.18 J82
 Effects of sericea lespedeza residues on warm-season grasses.
 Kalburtji, K.L.; Mosjidis, J.A.
 Denver, Colo. : Society for Range Management; 1992 Sep.
 Journal of range management v. 45 (5): p. 441-444; 1992 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lespedeza cuneata; Litter (plant); Plant
 residues; Allelopathy; Cynodon dactylon; Paspalum notatum;
 Seed germination; Seedling emergence; Inhibition; Cultivars;
 Varietal susceptibility; Tolerance; Nitrogen fertilizers
 
 Abstract:  Soil incorporation of sericea lespedeza [Lespedeza
 cuneata (Dum. de Cours) G. Don.] residues has been reported to
 inhibit growth of some forage grasses. No information is
 available on the performance of sericea lespedeza grown in
 association with warm-season perennial grasses. Laboratory and
 greenhouse experiments were conducted to determine if sericea
 lespedeza residues affect seed germination and seedling growth
 of bermudagrass [Cynodon dactylon (L.) Pers.] and bahiagrass
 (Paspalum notatum Flugge); if any such response was cultivar
 dependent; and if the response was subject to manipulation by
 N fertilization. Sericea lespedeza residues inhibited
 bermudagrass and bahiagrass growth, but did not affect their
 seed germination and emergence. No differences among cultivars
 of bermudagrass and bahiagrass in response to sericea
 lespedeza residues were found in the greenhouse. Nevertheless,
 differences among bermudagrass cultivars for tolerance to
 sericea lespedeza residues were observed in the laboratory.
 The harmful effects of sericea lespedeza residues were small
 (17 and 16% reduction of dry weight for bermudagrass and
 bahiagrass, respectively) compared to the positive effects of
 N fertilization.
 
 
 126                                NAL Call. No.: QD415.A1J6
 Effects of soil nitrogen level on ferulic acid inhibition of
 cucumber leaf expansion.
 Klein, K.; Blum, U.
 New York, N.Y. : Plenum Press; 1990 Apr.
 Journal of chemical ecology v. 16 (4): p. 1371-1383; 1990 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Leaves; Allelopathy; Ferulic
 acid; Soil chemistry; Nitrogen content; Seedlings; Growth
 
 Abstract:  It has been suggested that the allelopathic
 activity of phenolic acids should be primarily important in
 soils of low fertility. if this is true, then plant growth
 inhibition by phenolic acids may be unimportant in managed
 agricultural soils. The objective of this study was to
 determine how soil nitrogen (N) level might modify phenolic
 acid inhibition of growth. Cucumber seedlings (Cucumis sativus
 cv Early Green Cluster) grown in containers in growth chambers
 under varying N levels (5, 10, 15, 20, and 25 micrograms N/g
 soil) in Portsmouth B1-horizon soil material were treated with
 ferulic acid (0 or 10 micrograms/g soil). Nitrogen and ferulic
 acid (FA) were applied every other day to the soil surface.
 The amount of FA in the soil solution declined with depth in
 the containers. A more rapid disappearance of FA from the soil
 solution was observed for the last FA treatment (O% recovered
 after 10 hr on day 23) than the first treatment (44% recovered
 after 10 hr on day 13). Both low N (5 micrograms N/g soil) and
 FA treatments reduced shoot dry weight, the mean absolute
 (AGR) and the mean relative (RGR) rates of leaf expansion, and
 increased the root-shoot ratio. High N treatments reduced
 shoot dry weight and the AGR. Ferulic acid inhibited cucumber
 seedling growth over a range of N concentrations, suggesting
 that the allelopathic activity of phenolic acids may be
 important in both nutrient limiting and nonlimiting soils for
 some species.
 
 
 127                                NAL Call. No.: QD415.A1J6
 Effects of some compounds isolated from Celaenodendron
 mexicanum Standl (Euphorbiaceae) on seeds and phytopathogenic
 fungi.
 Castaneda, P.; Garcia, M.R.; Hernandez, B.E.; Torres, B.A.;
 Anaya, A.L.; Mata, R.; Effects of some compounds isolated from
 Celaenodendron mexicanum Standl (Euphorbiaceae) on seeds and
 phytopathogenic fun
 New York, N.Y. : Plenum Publishing Corporation; 1992 Jul.
 Journal of chemical ecology v. 18 (7): p. 1025-1037; 1992 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Mexico; Cabt; Euphorbiaceae; Plant composition;
 Flavones; Triterpenoids; Structure; Spectral data; Bioassays;
 Allelopathy; Seeds; Plant pathogenic fungi
 
 Abstract:  The known compounds friedelin, maytensifolin B,
 ginkgetin, bilobetin, and amentoflavone as well as the new
 triterpene 3 beta-hydroxyfriedelan-16-one were isolated from
 Celaenodendron mexicanum (Euphorbiaceae), an endemic tree of
 the Pacific Ocean coast of Mexico. The compounds' structures
 were established on the basis of spectral analysis. The
 biological effects of aqueous leachates, a CHCl3-MeOH extract
 and the isolated compounds of leaves and twigs were evaluated
 on the radicle growth of two plants, Amaranthus leucocarpus
 and Echinchloa crusgalli, and on the radial growth of three
 phytopathogenic fungi, Fusarium sp., Helminthosporium sp., and
 Alternaria sp. The organic extracts of both leaves and twigs
 inhibited Amaranthus and stimulated Echinchloa radicle growth.
 On the contrary, friedelin and maytensifolin B stimulated
 Amaranthus and inhibited Echinochloa. The target fungi showed
 a different response to each treatment, from inhibition to
 stimulation.
 
 
 128                                  NAL Call. No.: QL461.G4
 Effects of the allelochemical, alpha-tomatine, on the soybean
 looper (Lepidoptera: Noctuidae).
 Gallardo, F.; Boethel, D.J.
 Tifton, Ga. : Georgia Entomological Society; 1990 Jul.
 Journal of entomological science v. 25 (3): p. 376-382; 1990
 Jul.  Includes references.
 
 Language:  English
 
 Descriptors: Lycopersicon esculentum; Pest resistance;
 Pseudoplusia includens; Allelopathins; Alpha-tomatine; Diets;
 Larvae; Pupae; Weight gain; Maturation period; Survival;
 Growth rate
 
 
 129                                NAL Call. No.: QD415.A1J6
 Evaluation of flavonoids in Gossypium arboreum (L.) cottons as
 potential source of resistance to tobacco budworm.
 Hedin, P.A.; Jenkins, J.N.; Parrott, W.L.
 New York, N.Y. : Plenum Press; 1992 Feb.
 Journal of chemical ecology v. 18 (2): p. 105-114; 1992 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Gossypium arboreum; Heliothis virescens; Pest
 resistance; Flavonoids; Plant composition; Allelochemicals;
 Insect control; Biological control
 
 Abstract:  Asiatic cottons [Gossypium arboreum (L.)] have been
 investigated as a source of resistance to the tobacco budworm
 [Heliothis virescens (Fab.)] because their diversely colored
 petals have been presumed to contain various allelochemicals.
 However, we found that larvae fed G. arboreum squares (buds)
 grew about equally compared with those fed squares from
 commercial G. hirsutum lines. The best source of resistance
 was found in several G. hirsutum double-haploid (DH) lines. In
 our investigation of allelochemicals, G. arboreum lines were
 found to contain much less gossypol in leaves, squares (buds),
 and petals than G. hirsutum L. lines. Flavonoids were
 significantly higher in G. arboreum lines only in petals. Of
 22 G. arboreum lines from which squares were gathered and fed
 to tobacco budworm (TBW) larvae in the laboratory, larval
 growth was not significantly decreased on any, but larval
 survival was decreased on six. When the square flavonoids were
 isolated and incorporated in laboratory diets for the TBW,
 moderate toxicity was observed. However, the estimated
 toxicities were not greater than those of the same flavonoid
 isolates from G. hirsutum lines. The most prevalent
 flavonoids, all previously found in G. arboreum plant tissues,
 were gossypetin 8-0-glucoside and gossypetin 8-0-rhamnoside,
 neither of which were present in G. hirsutum tissue. Quercetin
 3-0-glucoside, quercetin-3'-0-glucoside, and quercetin 7-0-
 glucoside were also present in significant amounts in both
 species. Gossypetin 8-0-rhamnoside and gossypetin 8-0-
 glucoside were the most toxic flavonoids tested (the ED50% was
 estimated to be 0.007 and 0.024) and therefore may prove to be
 contributing factors of resistance to TBW feeding.
 
 
 130                                NAL Call. No.: QD415.A1J6
 Evidence for allelopathy by tree-of-heaven (Ailanthus
 altissima). Heisey, R.M.
 New York, N.Y. : Plenum Press; 1990 Jun.
 Journal of chemical ecology v. 16 (6): p. 2039-2055; 1990 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Ailanthus altissima; Allelopathy; Root exudates;
 Phytotoxicity; Succession
 
 Abstract:  Ailanthus altissima (Mill.) Swingle contains one or
 more phytotoxic compounds in roots and leaves. Activity is
 higher in roots, where it occurs primarily in the bark.
 Powdered root bark and lealfets strongly inhibited growth of
 garden cress (Lepidium sativum L.) when mixed with soil in
 Petri dishes (ID50 values = 0.03 g root bark, 0.6 g
 leaflet/dish). The toxic material was readily extracted by
 methanol but not dichloromethane. Pieces of root bark mixed
 with soil at 2, 1, and 0.5 g/pot reduced cress biomass in the
 greenhouse, whereas methanol-extracted root bark did not. The
 inhibitory effect of Ailanthus tissues in soil was short-lived
 (less than or equal to 4 weeks in pots in greenhouse, less
 than or equal to 3 days in Petri dishes in laboratory).
 Inhibition by root bark was sometimes superseded by
 stimulation. Fresh Ailanthus root segments placed in or on
 soil reduced growth of nearby cress seedlings. Fine roots were
 more inhibitory than coarse, and inhibition became more
 pronounced with increased time of soil exposure to roots. Soil
 collected near Ailanthus roots in the field supported reduced
 radicle growth of cress compared to control soil. In contrast,
 stemflow from Ailanthus trees stimulated cress growth. The
 results suggest allelopathy caused by toxin exudation from
 roots may contribute to the aggressiveness and persistence of
 Ailanthus in certain habitats.
 
 
 131                                  NAL Call. No.: 79.8 W41
 Evidence that sweet potato (Ipomoea batatas) is allelopathic
 to yellow nutsedge (Cyperus esculentus).
 Harrison, H.F. Jr; Peterson, J.K.
 Champaign, Ill. : Weed Science Society of America; 1991 Apr.
 Weed science v. 39 (2): p. 308-312; 1991 Apr.  Includes
 references.
 
 Language:  English
 
 Descriptors: South Carolina; Ipomoea batatas; Allelopathy;
 Cyperus esculentus; Weed control; Biological control;
 Competitive ability; Crop weed competition; Roots; Growth
 rate; Inhibition; Plant extracts; Periderm; Crop yield; Tubers
 
 Abstract:  In field studies, 'Regal' sweet potato greatly
 reduced yellow nutsedge growth when the two species were grown
 together using standard cultural practices. At the end of the
 growing season, yellow nutsedge shoot dry weight per m2, in
 plots where the two species were planted together was less
 than 10% of shoot weight in plots where nutsedge was grown
 alone. Presence of yellow nutsedge did not markedly affect
 sweet potato growth. When grown together in a greenhouse
 experiment designed to minimize the competitive effects of
 sweet potato on yellow nutsedge, yellow nutsedge growth was
 reduced more than 50% by sweet potato 8 and 12 weeks after
 planting. The most polar fraction of serially extracted sweet
 potato periderm tissue was highly inhibitory to yellow
 nutsedge root growth. These results indicate that sweet potato
 interference with yellow nutsedge under field conditions is
 partially due to allelopathy.
 
 
 132                                 NAL Call. No.: 60.18 J82
 Factors affecting weeping lovegrass seedling vigor on shinnery
 oak range. Matizha, W.; Dahl, B.E.
 Denver, Colo. : Society for Range Management; 1991 May.
 Journal of range management v. 44 (3): p. 223-227; 1991 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Texas; Eragrostis curvula; Seedlings; Vigor; Crop
 establishment; Quercus havardii; Rangelands; Environmental
 factors; Plant competition; Allelopathy; Cenchrus incertus;
 Plant residues; Nitrogen fertilizers; Seedbed preparation
 
 Abstract:  Low vigor of seedlings and stand failures plague
 many revegetation efforts in semiarid and arid rangelands.
 Phototoxicity, sandbur (Cenchrus incertus M.A. Curtis)
 competition, seedbed preparation (plowing vs. disking), and
 nitrogen (N) fertilization were studied as reasons for low
 vigor of Ermelo weeping lovegrass [Eragrostis curvula Schrad.)
 Nees] seedlings on sand shinnery oak (Quercus havardii Rydb.)
 range in west Texas. Oak leaf residue and sandbur-dominated
 grass residue extracts did not affect seed germination and
 initial shoot growth of lovegrass seedlings. However, these
 residue extracts reduced root length 92% and 21%,
 respectively. Survival of weeping lovegrass seedlings was not
 affected by even 65 sandbur plants/m2. But, herbage yield was
 reduced 65, 72, and 79% with 30, 45, and 65 sandbur plants/m2.
 Early in the growing season, unfertilized plowed (P) plots had
 5.6 ppm N in the 10-20 cm soil layer compared to a maximum of
 3.9 ppm on other seedbed treatments. In the surface 10 cm, the
 P plots had less N than the disked plots. Surface-applied N
 fertilizer accumulated in the upper 10 cm of soil and promoted
 weed growth without improving weeping lovegrass stands or
 seedling vigor. Weeping lovegrass seedling vigor was greatest
 on P and least on disked plots. Thus, plowing buried weed
 seeds better, put resident N more deeply into the soil for
 better root uptake, removed allelopathic residues from
 seedling contact better, and provided for much higher seedling
 vigor than the disked seedbeds.
 
 
 133                                 NAL Call. No.: QP33.J681
 Fate of plant-derived secondary metabolites in three moth
 species (Syntomis mogadorensis, Syntomeida epilais, and
 Creatonotos transiens). Wink, M.; Schneider, D.
 Berlin, W. Ger. : Springer International; 1990.
 Journal of comparative physiology : B : Biochemical,
 systematic, and environmental physiology v. 160 (4): p.
 389-400; 1990.  Includes references.
 
 Language:  English
 
 Descriptors: Nerium oleander; Plants; Lepidoptera; Larvae;
 Allelochemicals; Secondary metabolites; Alkaloids; Feeding
 preferences; Metabolism
 
 
 134                                 NAL Call. No.: 450 P5622
 Fatty acid incorporation in the biosynthesis of anacardic
 acids of geraniums. Walters, D.S.; Craig, R.; Mumma, R.O.
 Oxford : Pergamon Press; 1990.
 Phytochemistry v. 29 (6): p. 1815-1822. ill; 1990.  Includes
 references.
 
 Language:  English
 
 Descriptors: Pelargonium; Hybrids; Anacardic acid;
 Biosynthesis; Fatty acids; Trichomes; Chemical composition;
 Allelopathins; Insect pests; Insect control; Pest resistance;
 Radioactive tracers
 
 
 135                                  NAL Call. No.: 421 J822
 Feeding and growth responses of laboratory and field strains
 of velvetbean caterpillars (Lepidoptera: Noctuidae) to food
 nutrient level and allelochemicals.
 Slansky, F. Jr; Wheeler, G.S.
 Lanham, Md. : Entomological Society of America; 1992 Oct.
 Journal of economic entomology v. 85 (5): p. 1717-1730; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Anticarsia gemmatalis; Larvae; Diets; Feeding;
 Food consumption; Growth; Laboratory rearing; Responses
 
 Abstract:  The performance of Anticarsia gemmatalis (Hubner)
 larvae from a laboratory colony in existence for > 100
 generations was compared in several experiments with that of
 larvae from the field (first generation). In general, field-
 strain larvae exhibited prolonged development and a lower
 biomass-relative growth rate (RGR) when fed an artificial diet
 because of their slower biomass-relative consumption rate
 (RCR), compared with laboratory-strain larvae. In contrast,
 laboratory-strain larvae grew more slowly than field-strain
 larvae when both were fed foliage of the wild legume
 Indigofera hirsuta L., a larval foodplant of A. gemmatalis in
 the field. This slower growth rate was associated with a
 reduced efficiency of conversion of digested food to biomass
 (ECD); the RCR of laboratory- and field-strain larvae fed
 foliage did not differ significantly. When fed artificial
 diets with progressively reduced nutrient levels (as a
 percentage of fresh mass, fm), larvae of both strains
 compensated similarly by consuming more food (fm) at a faster
 rate (RCRfm). Nonetheless, total nutrient intake, biomass
 gain, pupal dry mass (percent fin), and pupal lipid content
 (percent dry mass, dm) for both strains declined on the diet
 with the lowest nutrient level. Nutrient utilization
 efficiencies also changed on the most diluted diet; for both
 strains, approximate digestibility of nutrients (ADnu)
 increased and ECDnu decreased. When fed artificial diet
 containing either the phenylcoumarin isoflavonoid, coumestrol
 (an 'evolutionarily familiar' allelochemical to A.
 gemmatalis), or the methylxanthine alkaloid, caffeine (an
 'evolutionarily novel' compound to this species),
 significantly higher mortality occurred among field-strain
 larvae. Whether this difference was caused by their inherently
 greater sensitivity to these allelochemicals compared with the
 laboratory-strain larvae, or to the synergistic effect of the
 additional stress imposed on them from consuming the
 artificial diet, was not det
 
 
 136                                 NAL Call. No.: 64.8 C883
 Field apparatus for testing allelopathy of annual bluegrass on
 creeping bentgrass.
 Brede, A.D.
 Madison, Wis. : Crop Science Society of America; 1991 Sep.
 Crop science v. 31 (5): p. 1372-1374; 1991 Sep.  Includes
 references.
 
 Language:  English
 
 Descriptors: Agrostis stolonifera var. palustris; Crop weed
 competition; Poa annua; Competitive ability; Allelopathy;
 Leachates; Field experimentation; Apparatus; Design; Golf
 green soils
 
 Abstract:  Golf-course superintendents have long observed the
 competitive nature of annual bluegrass (Poa annua L.) as a
 weed on creeping bentgrass [Agrostis stolonifera L. var.
 palustris (Huds.) Farw.] putting greens. Allelopathy has been
 suggested as a contributing factor in this competitiveness.
 This study tested the allelopathy hypothesis under putting-
 green conditions using a modified field approach of the
 conventional stair-step experimental procedure. Annual blue-
 grass and creeping bentgrass sand putting greens, each 297 m2,
 were established, and leachate from these greens was used to
 irrigate replicated sand-based creeping bentgrass test greens.
 Moisture sensing and irrigation of the test plots were under
 continuous electronic control. The leachate sampling and
 delivery system functioned flawlessly throughout the 2-yr
 period, in spite of weather extremes (>40 degrees C). After
 two growing seasons of monthly monitoring, no consistent
 effects on turf color, foliar ground cover, shoot density, or
 disease incidence were found in the test green to indicate
 allelopathy.
 
 
 137                                NAL Call. No.: QD415.A1J6
 Floral volatiles of Tanacetum vulgare L. attractive to Lobesia
 botrana Den. et Schiff. females.
 Gabel, B.; Thiery, D.; Suchy, V.; Marion-Poll, F.; Hradsky,
 P.; Farkas, P. New York, N.Y. : Plenum Press; 1992 May.
 Journal of chemical ecology v. 18 (5): p. 693-701; 1992 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Tanacetum vulgare; Lobesia botrana; Vitis
 vinifera; Flowers; Extracts; Plant composition;
 Allelochemicals; Terpenoids; Smell; Insect control
 
 Abstract:  The European grapevine moth (EGVM), Lobesia
 botrana, is a major pest of grapes in Europe. Females are
 attracted to a nonhost plant: tansy (Tanacetum vulgare L.),
 which is a common weed in Slovakian vineyards. A steam
 distillate extract of tansy flowers was analyzed by means of a
 GC-EAG technique to screen constituents detected by the
 olfactory receptors of EGVM females. From more than 200 GC
 peaks, nine peaks corresponding to monoterpenoids released an
 EAG response in more than 70% of the females (N = 15): p-
 cymene, d-limonene, alpha-thujene, alpha-thujone, beta-
 thujone, thujyl alcohol, terpinene-4-ol, (Z)-verbenol, and
 piperitone. The steam distillate of tansy as well as a
 synthetic blend of identified compounds released consistent
 attraction in a field cage. The use of nonhost plants and host
 plant odors in integrated pest management is discussed.
 
 
 138                            NAL Call. No.: QL542.C38 1993
 Foraging with finesse: caterpillar adaptations for
 circumventing plant defenses.
 Dussourd, D.E.
 New York : Chapman & Hall; 1993.
 Caterpillars : ecological and evolutionary constraints on
 foraging / edited by Nancy E. Stamp and Timothy M. Casey. p.
 92-131; 1993.  Includes references.
 
 Language:  English
 
 Descriptors: Lepidoptera; Feeding behavior; Adaptability;
 Plant composition; Defense mechanisms; Allelochemicals
 
 
 139                                NAL Call. No.: QD415.A1J6
 Formononetin 7-O-glucoside (ononin), an additional growth
 inhibitor in soils associated with the weed, Pluchea
 lanceolata (DC) C.B. Clarke (Asteraceae). Inderjit; Dakshini,
 K.M.M.
 New York, N.Y. : Plenum Press; 1992 May.
 Journal of chemical ecology v. 18 (5): p. 713-718; 1992 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Pluchea; Allelopathy; Phytotoxins; Soil analysis;
 Phenolic compounds; Bioassays; Seed germination; Flavonoids;
 Inhibition; Weed control
 
 Abstract:  Formononetin 7-O-glucoside (ononin), an
 isoflavonoid, was isolated from the soils of cultivated areas
 associated with the perennial weed, Pluchea lanceolata.
 Aqueous solutions of this compound inhibited significantly
 root and shoot growth of mustard at 1 X 10(-4) M, 5 X 10(-4)
 M, and 1 X 10(-3) M. The level of inhibition was similar to
 that of hesperidin and taxifolin 3-arabinoside, as reported
 earlier. The potential allelopathic effect of this compound
 under field conditions is discussed.
 
 
 140                                NAL Call. No.: QD415.A1B5
 Genetic variation in response of the gypsy moth to aspen
 phenolic glycosides. Lindroth, R.L.; Weisbrod, A.V.
 Oxford : Pergamon Press; 1991.
 Biochemical systematics and ecology v. 19 (2): p. 97-103;
 1991.  Includes references.
 
 Language:  English
 
 Descriptors: Lymantria dispar; Genetic variation; Populus;
 Phenolic compounds; Glycosides; Detoxification; Feeding;
 Esterases; Enzyme activity; Allelochemicals
 
 Abstract:  We investigated genetic variation in response of a
 wild strain of gypsy moths, Lymantria dispar, to phenolic
 glycosides isolated from Populus. For 12 families, first
 instar survival varied little among family groups fed a
 control artificial diet, but LT50, values differed
 significantly among groups fed an artificial diet containing
 phenolic glycosides. First instar survival rates were
 positively correlated with esterase enzyme activity. Relative
 growth rates of fourth instars were influenced by diet and by
 family. A nearly significant diet X family interaction term
 indicates that family (genotype) influences gypsy moth
 response to phenolic glycosides. Prior consumption of phenolic
 glycosides induced esterase activity in fifth instars.
 Esterase activity was influenced by diet, family and diet X
 family interaction. Genetic variation in esterase activity may
 explain variation in larval response to phenolic glycosides.
 
 
 141                                 NAL Call. No.: 60.18 J82
 Germination of 2 legumes in leachate from introduced grasses.
 Fulbright, N.; Fulbright, T.E.
 Denver, Colo. : Society for Range Management; 1990 Sep.
 Journal of range management v. 43 (5): p. 466-467; 1990 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: Dichanthium annulatum; Introduced species;
 Cenchrus ciliaris; Phytotoxins; Leachates; Germination
 inhibitors; Desmanthus; Cassia; Seed germination; Allelopathy;
 Phytotoxicity
 
 
 142                                 NAL Call. No.: QL496.J68
 Heritable differences in the response of the braconid wasp
 Microplitis croceipes to volatile allelochemicals.
 Prevost, G.; Lewis, W.J.
 New York, N.Y. : Plenum Publishing; 1990 May.
 Journal of insect behavior v. 3 (3): p. 277-287; 1990 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Glabromicroplitis croceipes; Parasitoids; Animal
 behavior; Heritability; Allelochemicals
 
 
 143                                NAL Call. No.: QD415.A1J6
 Hesperetin 7-rutinoside (Hesperidan) and taxifolin 3-
 arabinoside as germination and growth inhibitors in soils
 associated with the weed, Pluchea lanceolata (DC) C.B. Clarke
 (Asteraceae).
 Inderjit; Dakshini, K.M.M.
 New York, N.Y. : Plenum Press; 1991 Aug.
 Journal of chemical ecology v. 17 (8): p. 1585-1591; 1991 Aug. 
 Includes references.
 
 Language:  English
 
 Descriptors: Pluchea; Plant composition; Allelopathy; Seed
 germination; Bioassays; Extracts; Phenolic compounds; Weed
 control
 
 Abstract:  Hesperetin 7-rutinoside (Hesperidin) and taxifolin
 3-arabinoside were detected in the soils associated with the
 rapidly spreading perennial weed, Pluchea lanceolata. In the
 present investigations, inhibitory potential of the aqueous
 extracts of the two compounds was established and confirmed
 through growth experiments pertaining to seed germination and
 seedling growth of radish, mustard, and tomato, with 10(-4) M
 solutions or the authentic samples. The significance of the
 water-soluble compounds present in the rhizosphere zones of
 the weed and its interference potential is commented upon.
 
 
 144                                 NAL Call. No.: SB951.P49
 High-affinity juvenile hormone binding to fat body cytosolic
 proteins of the bollworm, Heliothis zea: characterization and
 interaction with allelochemicals and xenobiotics.
 Muehleisen, D.P.; Plapp, F.W. Jr; Benedict, J.H.; Carino, F.A.
 Duluth, Minn. : Academic Press; 1990 May.
 Pesticide biochemistry and physiology v. 37 (1): p. 64-73;
 1990 May.  Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Heliothis zea; Larvae; Fat body;
 Carrier proteins; Binding site; Juvenile hormones;
 Competition; Analogs; Displacement; Methoprene; Ocimene;
 Myrcene; Gossypol; Stereochemistry
 
 
 145                                   NAL Call. No.: QK1.A28
 In vitro inhibitory activity of some rhizosphere fungi of
 soybean against Sclerotium rolfsii SACC growth.
 Deb, P.R.
 Meerut, India : Society for Advancement of Botany; 1990 Dec.
 Acta botanica Indica v. 18 (2): p. 159-162; 1990 Dec. 
 Includes references.
 
 Language:  English
 
 Descriptors: Glycine max; Aspergillus; Fusarium; Penicillium;
 Trichoderma; Corticium rolfsii; Growth inhibitors; Culture
 filtrates; Rhizosphere fungi; Allelopathins; Metabolites
 
 
 146                                 NAL Call. No.: 464.8 P56
 In vitro reactions of Cladosporium caryigenum with pecan
 condensed tannins and isoquercitrin.
 Laird, D.W.; Graves, C.H. Jr; Hedin, P.A.
 St. Paul, Minn. : American Phytopathological Society; 1990
 Feb. Phytopathology v. 80 (2): p. 147-150; 1990 Feb.  Includes
 references.
 
 Language:  English
 
 Descriptors: Carya pecan; Leaves; Plant extracts; Tannins;
 Isoquercitrin; Allelopathins; In vitro; Cladosporium; Growth
 rate; Growth retardation; Disease resistance; Fungicidal
 properties; Bioassays
 
 Abstract:  In vitro growth of Cladosporium caryigenum, the
 pecan scab incitant, was significantly inhibited by condensed
 tannin and isoquercitrin, allelochemicals that had been
 extracted from freeze-dried pecan leaves (Carya illinoensis
 cv. Van Deman). Nine isolates of C. caryigenum varied in their
 responses to condensed tannin at a concentration of 4,000
 microgram/ml. Even the most tolerant isolate was inhibited at
 concentrations of 4,000 microgram/ml and above. Isoquercitrin
 at 4,000 microgram/ml was about two to four times more
 inhibitory than tannin to growth of three isolates of C.
 caryigenum, and differences were found in the tolerance of the
 three isolates to isoquercitrin.
 
 
 147                                 NAL Call. No.: 421 EN895
 In vivo effect of mixtures of allelochemicals on the life
 cycle of the European corn borer, Ostrinia nubilalis.
 Bernard, C.B.; Arnason, J.T.; Philogene, B.J.R.; Lam, J.;
 Waddell, T. Dordrecht : Kluwer Academic Publishers; 1990 Oct.
 Entomologia experimentalis et applicata v. 57 (1): p. 17-22;
 1990 Oct. Includes references.
 
 Language:  English
 
 Descriptors: Ostrinia nubilalis; Plant pests; Growth; Life
 cycle; Mortality; Toxicity; Allelochemicals; Compositae;
 Insecticidal action; Plant extracts
 
 
 148                                 NAL Call. No.: 421 EN895
 Induction of aldrin epoxidation and glutathione S-transferase
 in te mite Rhizoglyphus robini.
 Capua, S.; Cohen, E.; Gerson, U.
 Dordrecht : Kluwer Academic Publishers; 1991 Apr.
 Entomologia experimentalis et applicata v. 59 (1): p. 43-50;
 1991 Apr. Includes references.
 
 Language:  English
 
 Descriptors: Allium sativum; Daucus carota; Rhizoglyphus
 robini; Aldrin; Allelochemicals; Enzyme activity; Epoxides;
 Glutathione transferase; Pesticide resistance; Pesticides;
 Toxic substances
 
 
 149                                NAL Call. No.: QD415.A1J6
 Influence of Artemisia princeps var. orientalis components on
 callus induction and growth.
 Kil, B.S.; Yun, K.W.; Lee, S.Y.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Aug.
 Journal of chemical ecology v. 18 (8): p. 1455-1462; 1992 Aug. 
 Includes references.
 
 Language:  English
 
 Descriptors: Artemisia princeps; Allelopathy; Essential oils;
 Plant composition; Callus
 
 Abstract:  An in vitro study was performed to determine the
 potential application of tissue culture in determining
 allelopathic potential of Artemisia princeps var. orientalis
 (wormwood). Aqueous extracts and volatile substances of A.
 princeps var. orientalis were tested to determine their
 effects on callus induction and growth of several tested
 species. Extracts of 5% A. princeps var. orientalis caused
 some reduction in concentration, induction, and growth of
 callus, although they looked normal, whereas the explants of
 most receptor plants did not develop callus at higher
 concentration. Lettuce and Eclipta prostrata were the most
 sensitive species, and A. princeps var. orientalis was
 affected by its own extracts. The growth of calluses in MS 121
 medium treated with essential oil of A. princeps var.
 orientalis was inhibited, and the degree of inhibition was
 proportional to the concentration of the essential oil.
 
 
 150                                  NAL Call. No.: QL495.A7
 Influence of light on plant allelochemicals: a synergistic
 defense in higher plants.
 Downum, K.R.; Swain, L.A.; Faleiro, L.J.
 New York, N.Y. : Wiley-Liss; 1991.
 Archives of insect biochemistry and physiology v. 17 (4): p.
 201-211; 1991. Paper presented at a symposium on biochemical
 strategies of offense and defense at the plant-insect
 interface, 1989, San Antonio, Texas.  Includes references.
 
 Language:  English
 
 Descriptors: Costa Rica; Insect pests; Pest resistance;
 Phototoxins; Allelochemicals
 
 Abstract:  Plant phototoxins are broad-spectrum biocides which
 adversely affect an array of potential plant enemies,
 including among others disease-causing pathogens, nematodes,
 insect herbivores, and competing plant species. Thus far,
 plants which contain these broad-spectrum allelochemicals have
 been found to occur in open habitats (i.e., in full sunlight)
 where a defensive mechanism mediated by light would seem to
 operate most effectively. The levels of available light in
 shaded environments, although considerably lower than full sun
 (1-10% of full sun), are equivalent to the intensities of
 light used to kill phototoxin-treated insects in laboratory
 studies. This suggests that phototoxic reactions might mediate
 important organismal interactions in shaded environments as
 well. In this study, more than 230 Costa Rican rainforest
 plants were bioassayed for phototoxic metabolites in an effort
 to ascertain their prevalence among plants growing in moderate
 to extreme shade. Microbial bioassays, employing Bacillus
 cereus (a gram positive bacterium), Escherichia coli (a gram
 negative bacterium), and Saccharomyces cerevisiae (a yeast)
 were used to rapidly and sensitively indicate phototoxic
 action and potential for insecticidal action. Tissue extracts
 from 12 plant families tested positive for phototoxins. This
 is the first report of phototoxins occurring in eight of those
 families (Acanthaceae, Campanulaceae, Gesnariaceae,
 Loganiaceae, Malpigaceae, Phytolaccaceae, Piperaceae, and
 Sapotaceae). The presence of phototoxins in rainforest plants
 suggests that phototoxic plant allelochemicals may function as
 important defenses in low-light, as well as high-light,
 environments.
 
 
 151                               NAL Call. No.: QL391.N4J62
 Influence of nonhost plants on population decline of
 Rotylenchulus reniformis. Caswell, E.P.; DeFrank, J.; Apt,
 W.J.; Tang, C.S.
 Lake Alfred, Fla. : Society of Nematologists; 1991 Jan.
 Journal of nematology v. 23 (1): p. 91-98; 1991 Jan.  Includes
 references.
 
 Language:  English
 
 Descriptors: Hawaii; Rotylenchulus reniformis; Chloris gayana;
 Crotalaria juncea; Digitaria decumbens; Tagetes patula;
 Allelopathy; Chitin; Soil amendments; Nematode control;
 Population dynamics
 
 
 152                                 NAL Call. No.: 79.8 W412
 Influence of pasture grass and legume swards on seedling
 emergence and growth of Carduus nutans L. and Cirsium vulgare
 L.
 Wardle, D.A.; Rahman, A.
 Oxford : Blackwell Scientific Publications; 1992 Apr.
 Weed research v. 32 (2): p. 119-128; 1992 Apr.  Includes
 references.
 
 Language:  English
 
 Descriptors: Pastures; Dactylis glomerata; Phalaris aquatica;
 Lolium perenne; Bromus catharticus; Holcus lanatus; Festuca
 arundinacea; Medicago sativa; Trifolium pratense; Trifolium
 subterraneum; Trifolium repens; Crop weed competition; Carduus
 nutans; Cirsium vulgare; Seedling emergence; Growth rate;
 Inhibition; Weed control; Biological control; Allelopathy
 
 
 153                                NAL Call. No.: QD415.A1J6
 Influence of phenolic acids on microbial populations in the
 rhizosphere of cucumber.
 Shafer, S.R.; Blum, U.
 New York, N.Y. : Plenum Press; 1991 Feb.
 Journal of chemical ecology v. 17 (2): p. 369-389; 1991 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Rhizosphere; Allelopathy;
 Ferulic acid; Coumaric acids; Vanillic acid; Soil bacteria;
 Soil fungi
 
 Abstract:  Experiments were conducted to determine whether
 changes in soil microbial populations that occur in response
 to additions of certain allelopathic phenolic acids to bulk
 soil also occur in the rhizosphere. Cucumber seedlings were
 transplanted into cups containing a nutrient-enriched mixture
 of Portsmouth B, soil and sand and were watered five times
 (once every 48 hr) with aqueous solutions of ferulic, p-
 coumaric, or vanillic acid (each at 0, 0.25, or 0.50
 micromoles/g soil material). Nutrient solution was applied on
 alternate days. Leaf growth was suppressed by up to 42% by
 phenolic acids, but changes in root growth varied with the
 compound and concentration in solution. Significant increases
 (over 600% relative to controls) in populations of fast-
 growing bacteria in the rhizosphere were detected after two
 but not after five treatments, and increases (400% relative to
 controls) in numbers of fungal propagules were detected after
 five treatments. Such increases suggested that chronic
 exposure to a phenolic acid might result in high populations
 of rhizosphere microorganisms that could metabolize the
 compounds and thus alter observable responses by the plant. To
 test this, plants were watered repeatedly with a low-
 concentration solution of ferulic acid (chronic treatments,
 0.0 or 0.1 micromoles/g soil material in one experiment, 0.000
 or 0.025 micromoles/g soil material in a second) and then once
 with a high-concentration solution (acute treatment; 0.0, 0.5,
 or 1.0 micromoles/g soil material in the first experiment;
 0.000, 0.125, or 0.250 micromoles/g soil material in the
 second). Acute treatments and some chronic treatments
 suppressed leaf growth, but results were inconsistent for root
 growth. Acute treatments increased numbers of several types of
 bacteria in the rhizosphere but had inconsistent effects on
 fungi. Chronic treatments had no effect on numbers of bacteria
 or fungal propagules in the rhizosphere. Furthermore, chronic
 treatments did not alter responses of
 
 
 154                                   NAL Call. No.: 410 EC7
 Influence of plant allelochemicals on the tobacco hornworm and
 its parasitoid, Cotesia congregata.
 Barbosa, P.; Gross, P.; Kemper, J.
 Tempe, Ariz. : The Society; 1991 Oct.
 Ecology : a publication of the Ecological Society of America
 v. 72 (5): p. 1567-1575; 1991 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Nicotiana; Alkaloids; Allelochemicals;
 Antifeedants; Hordenine; Nicotine; Rutoside; Manduca sexta;
 Biological control agents; Cotesia; Host parasite
 relationships; Mortality; Parasites of insect pests
 
 
 155                                NAL Call. No.: QD415.A1J6
 Influence of sagebrush terpenoids on mule deer preference.
 Bray, R.O.; Wambolt, C.L.; Kelsey, R.G.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2053-2062; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Odocoileus hemionus; Feeding behavior; Artemisia
 tridentata; Plant composition; Terpenoids; Allelochemicals;
 Palatability
 
 Abstract:  The effect on mule deer (Odocoileus hemionus
 Rafinesque) preference of compounds in mountain big sagebrush
 [Artemisia tridentata Nutt. ssp. vaseyana (Rydb.) Beetle],
 Wyoming big sagebrush (A. t. ssp. wyomingensis Beetle and
 Young), basin big sagebrush (A. t. ssp. tridentata), and black
 sagebrush (A. nova Nels.) was compared using a two-choice
 preference test. Compounds tested included: p-cymene, 1,8-
 cineole, methacrolein (two concentrations), and the
 nonvolatile crude terpenoid fraction (NVCTF) from each taxon.
 The compounds were tested by applying them to chopped alfalfa
 hay at concentrations similar to those found in nature. The
 intake of the treated hay was compared with that of an
 untreated control. Eight deer were used as test animals in an
 8 X 8 Latin-square design. All compounds tested significantly
 deterred ingestion (P < 0.05). Compound influence on
 preference, in order of increasing deterrence, was as follows:
 50% methacrolein, mountain big sagebrush NVCTF, methacrolein,
 basin big sagebrush NVCTF, p-cymene, Wyoming big sagebrush
 NVCTF, black sagebrush NVCTF, and 1,8-cineole. Methacrolein
 appears to be an important preference determinant among big
 sagebrush subspecies, and p-cymene between black sagebrush and
 big sagebrush. The NVCTFs containing sesquiterpene lactones as
 one of their constituents were closely related to the
 preference of all four taxa. Future studies of animal
 preference for sagebrush should consider all of the potential
 defensive chemicals in the foliage.
 
 
 156                                  NAL Call. No.: 79.8 W41
 Influence of tillage, crop rotation, and weed management on
 giant foxtail (Setaria faberi) population dynamics and corn
 yield.
 Schreiber, M.M.
 Champaign, Ill. : Weed Science Society of America; 1992.
 Weed science v. 40 (4): p. 645-653; 1992.  Paper presented at
 the "Symposium on crop/weed management and the dynamics of
 weed seedbanks," February 11, 1992, Orlando, Florida. 
 Includes references.
 
 Language:  English
 
 Descriptors: Indiana; Zea mays; Setaria faberi; Weed biology;
 Seed banks; Population density; Population dynamics; Plowing;
 No-tillage; Rotations; Allelopathy; Cropping systems; Crop
 yield; Weed control; Chemical control; Herbicides
 
 Abstract:  A long-term integrated pest management study
 initiated in 1980 and continued through 1991 was conducted to
 determine interactions of tillage, crop rotation, and
 herbicide use levels on weed seed populations, weed
 populations, and crop yield. This paper presents giant foxtail
 seed population and stand along with corn yield in continuous
 corn, corn rotated with soybean, or corn following wheat in a
 soybean-wheat-corn rotation. Increasing herbicide use levels
 above the minimum reduced giant foxtail seed in the 0- to 2.5-
 cm depth of soil. Reducing tillage from conventional moldboard
 plowing to chiseling to no-tilling increased giant foxtail
 seed in only the top 0 to 2.5 cm of soil. No-tilling increased
 giant foxtail seed over conventional tillage in each year data
 were collected. Growing corn in a soybean-corn or soybean-
 wheat-corn rotation reduced giant foxtail seed from corn grown
 continuously in all three soil depths sampled: 0 to 2.5 cm,
 2.5 to 10 cm, and 10 to 20 cm. Although stands of giant
 foxtail tended to follow soil weed seed counts, crop rotation
 significantly reduced giant foxtail stand with maximum
 reduction in the soybean-wheat-corn rotation in all tillage
 systems. Giant foxtail stands were reduced following wheat in
 no-tilling, probably because of the allelopathic influence of
 wheat straw. Corn yields showed weed management levels above
 minimum control are not justified regardless of tillage and
 crop rotation.
 
 
 157                                 NAL Call. No.: QL461.A52
 Ingested allelochemicals insect wonderland: a menu of
 remarkable functions. Blum, M.S.
 Lanham, Md. : Entomological Society of America; 1992.
 American entomologist v. 38 (4): p. 222-234. ill; 1992. 
 Includes references.
 
 Language:  English
 
 Descriptors: Insect pests; Plant pests; Allelochemicals;
 Antifeedants; Detoxification; Ingestion; Metabolism; Plant
 protection; Predators of insect pests
 
 
 158                                NAL Call. No.: QD415.A1J6
 Inhibition of cucumber leaf expansion by ferulic acid in
 split-root experiments.
 Klein, K.; Blum, U.
 New York, N.Y. : Plenum Press; 1990 Feb.
 Journal of chemical ecology v. 16 (2): p. 455-463; 1990 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Ferulic acid; Root treatment;
 Allelopathy; Leaf area; Growth analysis
 
 Abstract:  Experiments were conducted to determine how the
 proportion of a root system in contact with an allelopathic
 compound may affect seedling responses. Cucumber seedlings
 grown in a split-root nutrient culture system were given
 either single (1 mM) or multiple applications (0.5 mM) of
 ferulic acid. Seedlings receiving single applications were
 left in the treatment solutions for two days and then
 harvested, while seedlings receiving multiple applications had
 their solutions changed every other day for a total of three
 changes. Leaf areas were determined daily starting with the
 initial ferulic acid treatment. Mean absolute and mean
 relative rates of leaf expansion were inversely related to the
 proportion of the root system in ferulic acid solution. Leaf
 expansion was inhibited primarily during the first 24 hr after
 each treatment. A partial recovery of growth occurred during
 the second 24-hr period following each treatment. Root length
 was reduced by ferulic acid. These results suggest that
 information on root and allelochemical distribution in soils
 is important when assessing the potential of allelopathic
 interactions between plants.
 
 
 159                                NAL Call. No.: QD415.A1J6
 Inhibition of Schizachyrium scoparium (Poaceae) by the
 allelochemical hydrocinnamic acid.
 Williamson, G.B.; Obee, E.M.; Weidenhamer, J.D.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Nov.
 Journal of chemical ecology v. 18 (11): p. 2095-2105; 1992
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Schizachyrium scoparium; Plant composition;
 Allelopathy; Cinnamic acid; Derivatives; Nitrogen; Phosphorus;
 Potassium; Scrub; Sandy soils; Bioassays; Growth; Inhibition;
 Greenhouse culture
 
 Abstract:  Bare zones around shrubs in the Florida scrub
 indicate the possibility of allelopathy by shrubs controlling
 the distribution of grasses invading from adjacent sandhills.
 The allelochemical, hydrocinnamic acid, has been identified as
 a breakdown product of ceratiolin, which is released from the
 shrub Ceratiola ericoides. Here, hydrocinnamic acid (HCA) was
 shown to have a strongly inhibitory effect on shoot and root
 biomass of the grass Schizachyrium scoparium in greenhouse
 bioassays lasting 4.5 months. Linear increases in the
 concentration of HCA from 0 to 200 ppm, applied biweekly,
 resulted in exponential decreases in root and shoot biomass at
 harvest. Plants grown at 200 ppm HCA had root and shoot
 biomasses 13% and 17% of controls, respectively. Concurrent
 investigation of reduced nutrient levels indicated greater
 inhibition by HCA in a reduced nitrogen (N) treatment and in a
 reduced potassium (K) treatment relative to HCA inhibition in
 the full nutrient treatment. The negative slopes of the
 regressions of log of biomass on HCA concentration were
 steepest in the reduced N and reduced K treatments. Root and
 shoot biomasses in reduced N treatments were 20-43% and 24-34%
 less than the respective biomasses in the full nutrient
 treatment. Comparable reductions in the reduced K treatment
 were as much as 19% and 10% for root and shoot biomasses,
 respectively. The effects of HCA in a reduced phosphorus (P)
 treatment and in a reduced P and K treatment were not
 significantly different from the effects of HCA in the full
 nutrient treatment. Extraction of the soils at harvest
 indicated no buildup of HCA at the end of the experiment. The
 sensitivity of Schizachyrium scoparium to HCA in general and
 increased sensitivity under low N and low K solutions may be
 important in the Florida scrub community where levels of N and
 K are known to be low.
 
 
 160                                NAL Call. No.: QD415.A1J6
 Inhibition of Scots pine seedling establishment by Empetrum
 hermaphroditum. Nilsson, M.C.; Zackrisson, O.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Oct.
 Journal of chemical ecology v. 18 (10): p. 1857-1870; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Pinus sylvestris; Empetrum; Leaves; Allelopathy;
 Phytotoxicity; Seedlings; Growth; Regeneration
 
 Abstract:  Poor establishment and reduced seedling growth of
 Scots pine (Pinus silvestris L.) in northern Sweden is related
 to an allelopathic inhibition by the dwarf shrub Empetrum
 hermaphroditum Hagerup. Indoor bioassays with green and brown
 leaves of Empetrum have strong negative effects on rooting
 ability radicle elongation, and growth of Scots pine
 seedlings. Bioassays with soil samples show that phytotoxic
 substances leached from Empetrum foliage accumulate in the
 soil. Field experiments reveal that chemical inhibition by
 Empetrum, causing high mortality and slow growth of pine
 seedlings, can be reduced by adding activated carbon to the
 soil.
 
 
 161                                 NAL Call. No.: 450 P5622
 Insect allelochemicals from Ajuga plants.
 Camps, F.; Coll, J.
 Oxford : Pergamon Press; 1993 Apr.
 Phytochemistry v. 32 (6): p. 1361-1370; 1993 Apr.  Literature
 review. Includes references.
 
 Language:  English
 
 Descriptors: Ajuga; Plant composition; Diterpenes; Molting
 hormones; Allelochemicals; Antifeedants; Plant products;
 Insect pests; Insect control; Literature reviews; Molecular
 conformation
 
 Abstract:  Clerodane diterpenoids and phytoecdysteroids with
 potential insect antifeedant and moulting hormone activities,
 respectively, have been isolated from Ajuga plants. Some
 clerodanes were active against larvae of Egyptian cotton
 leafworm, Spodoptera littoralis, when present in the diet at 3
 ppm doses. Structure-antifeedant activity relations were
 established. Likewise, first stage larvae of the greenhouse
 whitefly, Trialeurodes vaporariorum, exhibited complete
 mortality when fed on A. reptans. This effect was mainly
 originated by 29-norsengosterone and ajugalactone, two
 phytoecdysteroids occurring in this plant. For
 biotechnological production of phytoecdysteroids its total
 content in different parts of normally grown or in vitro
 micropropagated A. reptans plants was studied. Great
 quantitative and qualitative differences were observed. For
 comparison of these qualitative differences, a dealkylation
 ratio (Dr = C28/C29 phytoecdysteroid content) and a C-5
 hydroxylation ratio (5Hr = 5-OH/5-H phytoecdysteroid content)
 were established. The 5Hr values appeared to be quite constant
 ranging from 0.2 to 0.4, whereas Dr values oscillated from 2.3
 in whole plants to 12 in root cultures. Production of
 phytoecdysteroids was highest (approximately equal to 5000
 ppm/dry wt) in cultures of roots in an hormone supplemented
 solid medium.
 
 
 162                                  NAL Call. No.: QR53.J68
 Insect fungal symbionts: a promising source of detoxifying
 enzymes. Dowd, P.F.
 Amsterdam, The Netherlands : Published by Elsevier Science
 Publishers on behalf of the Society for Industrial
 Microbiology, c1986-; 1992 May. Journal of industrial
 microbiology v. 9 (3/4): p. 149-161; 1992 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Fungi; Mycotoxins; Secondary metabolites;
 Allelochemicals; Insecticides; Literature reviews
 
 Abstract:  Many species of insects cultivate, inoculate, or
 contain symbiotic fungi. Insects feed on plant materials that
 contain plant-produced defensive toxins, or are exposed to
 insecticides or other pesticides when they become economically
 important pests. Therefore, it is likely that the symbiotic
 fungi are also exposed to these toxins and may actually
 contribute to detoxification of these compounds. Fungi
 associated with bark beetles, ambrosia beetles, termites,
 leaf-cutting ants, long-horned beetles, wood wasps, and drug
 store beetles can variously metabolize/detoxify tannins,
 lignins, terpenes, esters, chlorinated hydrocarbons, and other
 toxins. The fungi (Attamyces) cultivated by the ants and the
 yeast (Symbiotaphrina) contained in the cigarette beetle gut
 appear to have broad-spectrum detoxifying abilities. The
 present limiting factor for using many of these fungi for
 large scale detoxification of, for example, contaminated soils
 or agricultural commodities is their slow growth rate, but
 conventional strain selection techniques or biotechnological
 approaches should overcome this problem.
 
 
 163                                  NAL Call. No.: QP501.C6
 Insect glutathione-S-transferase: a predictor of
 allelochemical and oxidative stress.
 Weinhold, L.C.; Ahmad, S.; Pardini, R.S.
 Oxford : Pergamon Press; 1990.
 Comparative biochemistry and physiology : B : Comparative
 biochemistry v. 95 (2): p. 355-363; 1990.  Includes
 references.
 
 Language:  English
 
 Descriptors: Spodoptera eridania; Papilio; Glutathione;
 Transferases; Stress factors; Enzyme activity
 
 
 164                                  NAL Call. No.: 79.8 W41
 Interaction of light, soil moisture, and temperature with weed
 suppression by hairy vetch residue.
 Teasdale, J.R.
 Champaign, Ill. : Weed Science Society of America; 1993 Jan.
 Weed science v. 41 (1): p. 46-51; 1993 Jan.  Includes
 references.
 
 Language:  English
 
 Descriptors: Vicia; Vetch; Cover crops; Light relations;
 Allelopathy; Soil water; Temperature; Shade; Establishment;
 Zea mays; Abutilon theophrasti; Setaria viridis; Chenopodium
 album; Night temperature; Weed control; Suppression
 
 Abstract:  The influence of light, soil moisture. and
 temperature on establishment of selected species through hairy
 vetch residue on the soil surface was investigated under
 controlled conditions in the greenhouse. Hairy vetch residue
 at rates ranging from 0 to 616 g m-2 had no effect on corn,
 slightly reduced velvetleaf and green foxtail establishment,
 and severely inhibited common lambsquarters establishment
 under full sunlight conditions. The same rates of hairy vetch
 residue reduced velvetleaf, green foxtail, and common
 lambsquarters establishment more under a shade cloth with 9%
 light transmittance than under full sunlight. Day/night
 temperatures of 24/16 or 32/26 degrees C had no effect and
 soil moistures of 50 or 133% field capacity had little effect
 on response of all species to residue rates. Weed
 establishment was similar under shade cloth without residue as
 under residue with an equivalent light transmittance,
 suggesting that light was more important than allelopathy or
 physical impedance for weed suppression by hairy vetch
 residue.
 
 
 165                                  NAL Call. No.: 450 AM36
 Interference potential of Pluchea lanceolata (Asteraceae):
 growth and physiological responses of asparagus bean, Vigna
 unguiculata var. sesquipedalis.
 Inderjit; Dakshini, K.M.M.
 Columbus, Ohio : Botanical Society of America; 1992 Sep.
 American journal of botany v. 79 (9): p. 977-981; 1992 Sep. 
 Includes references.
 
 Language:  English
 
 Descriptors: India; Vigna unguiculata subsp. sesquipedalis;
 Crop weed competition; Germination inhibitors; Growth
 inhibitors; Plant physiology; Pluchea; Allelochemicals;
 Allelopathy; Leachates; Plant extracts; Soil properties
 
 Abstract:  The water-soluble compounds synthesized by the
 weed, Pluchea lanceolata, and released by it into the soil
 significantly reduced seed germination, number of nodes,
 internode length, shoot and root lengths, nodule number and
 weight, and Chl a and b and Chl a/b ratio of asparagus bean
 plants. The pattern of accumulation of nutrients in shoot and
 root of asparagus bean was also affected. In contrast, the net
 photosynthetic rate and stomatal conductance of fully expanded
 leaves were higher in plants grown with treated soil. The
 concentrations of Mg++, Zn++, and PO4(3-) were higher and K+
 was lower in shoots of plants grown with treated soil as
 compared to those grown with the control soil. Also, roots of
 plants grown with treated soil showed greater accumulation of
 Mg++ and NO3-. Shoot/root ratio of nutrients in plants grown
 with control soil were higher for Zn++, Na+, Ca++, and NO3-,
 whereas plants grown with treated soil had higher ratios for
 PO4(3-). These results provide evidence for allelopathic
 interference by P. lanceolata to the growth of asparagus bean.
 
 
 166                                NAL Call. No.: QD415.A1J6
 Investigations on some aspects of chemical ecology of
 cogongrass, Imperata cylindrica (L.) Beauv.
 Inderjit; Dakshini, K.M.M.
 New York, N.Y. : Plenum Press; 1991 Feb.
 Journal of chemical ecology v. 17 (2): p. 343-352; 1991 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Imperata cylindrica; Allelopathy; Rhizomes;
 Leaves; Soil; Leachates; Phenolic compounds; Inhibition; Seed
 germination; Root nodules; Soil fungi; Melilotus indica
 
 Abstract:  To understand the interference mechanism of the
 weed, cogongrass, Imperata cylindrica (L.) Beauv., its effect
 on nutrient availability and mycoflora of its soil rhizosphere
 as well as nodule characteristics, root length, and root/shoot
 ratio of Melilotus parviflora Desf. were investigated.
 Additionally, the effect of the leachates of leaves and
 root/rhizome of cogongrass on seed germination and seedling
 characteristics of radish, mustard, fenugreek, and tomato were
 examined. Furthermore, to assess the qualitative and
 quantitative differences in phytochemical components, the
 leachates and the soils from three sampling sites (with
 cogongrass and 1.5 m and 3 m away from cogongrass) were
 analyzed with high-performance liquid chromatography (HPLC) on
 a C18 column. No significant difference in nutrient
 availability was found, but qualitative and quantitative
 differences in phenolic fractions were recorded in the three
 sampling sites. Furthermore, of the 19 fungi recorded in the
 soils, decreases in the number of colonies (per gram of soil)
 of Aspergillus fumigatus, A. niger, A. candidus, and an
 increase of A. flavus was recorded in the soils with
 cogongrass. The inhibition in nodule number, weight, nitrogen
 fixation (acetylene reduction activity), root length, and
 root/shoot ratio of Melilotus parviflora were noted. Percent
 seed germination, root and shoot length, fresh and dry weight
 of seedlings of different seeds were affected by the leachates
 of leaves and root/rhizome. It was found that root/rhizome
 leachate was more inhibitory than leaf leachate. However, the
 inhibition was higher in soil + leaves leachate than soil +
 root/rhizome leachate. HPLC analysis established that four
 compounds were contributed by the weed to the soil system even
 though their relative concentration varies in various
 leachates. It is surmised that these compounds cause
 allelopathic inhibition of growth characteristics of seeds
 tested. Significance of the data vis-a-vis the interference
 poten
 
 
 167                                  NAL Call. No.: 500 N21P
 Involvement of cytochrome P450 in host-plant utilization by
 Sonoran Desert Drosophila.
 Frank, M.R.; Fogleman, J.C.
 Washington, D.C. : The Academy; 1992 Dec15.
 Proceedings of the National Academy of Sciences of the United
 States of America v. 89 (24): p. 11998-12002; 1992 Dec15. 
 Includes references.
 
 Language:  English
 
 Descriptors: Host plants; Cytochrome p-450; Drosophila;
 Alkaloids; Allelochemicals; Ecosystems; Larvae; Metabolism;
 Oxygenases; Vigor
 
 Abstract:  The four Drosophila species endemic to the Sonoran
 Desert (Drosophila mettleri, Drosophila mojavensis, Drosophila
 nigrospiracula, and Drosophila pachea) utilize necrotic cactus
 tissue or soil soaked by rot exudate as breeding substrates.
 Each Drosophila species uses a different cactus species as its
 primary host. D. pachea is limited to senita cactus by a
 biochemical dependency on unusual sterols available only in
 that cactus. For the other Drosophila species, no such
 chemical dependencies exist to explain the relationships with
 their primary host plants. Each cactus species has a different
 array of allelochemicals that have detrimental effects on
 nonresident fly species. We have hypothesized that the desert
 fly-cactus associations are due, in part, to differences
 between the fly species in their allelochemical detoxication
 enzymes, the cytochrome P450 system. To test whether P450s are
 involved in the detoxication of cactus allelochemicals,
 several experiments were done. (i) The effect of a specific
 P450 inhibitor, piperonyl butoxide, on larval survival through
 eclosion on each cactus substrate was investigated. (ii) In
 vitro metabolism of cactus alkaloids was determined for each
 Drosophila species. The effects of specific inducers and
 inhibitors were included in these experiments. (iii) The basal
 and induced content of cytochrome P450 in each species was
 determined. The results support the hypothesis that P450
 enzymes are involved in host-plant utilization by these
 Sonoran Desert Drosophila species.
 
 
 168                                NAL Call. No.: QD415.A1J6
 Isolation and characterization of phytotoxic compounds from
 asparagus (Asparagus officinalis L.) roots.
 Hartung, A.C.; Nair, M.G.; Putnam, A.R.
 New York, N.Y. : Plenum Press; 1990 May.
 Journal of chemical ecology v. 16 (5): p. 1707-1718; 1990 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Asparagus officinalis; Roots; Allelopathy;
 Phytotoxicity; Extracts; Chemical composition; Bioassays;
 Lepidium sativum
 
 Abstract:  Potential allelochemicals from aqueous extracts of
 dried asparagus (Asparagus officinalis L.) roots were isolated
 and characterized. Active fractions separated by HPLC included
 ferulic, isoferulic, malic, citric, and fumaric acids. Soxhlet
 extraction of the residues also produced phytotoxic caffeic
 acid. Although none of these compounds, when applied singly,
 was active enough to account for the phytotoxicity of
 asparagus extracts, their combined effect might be additive or
 synergistic. An extract from lyophilized fresh root tissues
 contained a fraction that was one order of magnitude more
 toxic than any compound obtained from the dried roots. The
 most active component was isolated by TLC and characterized by
 [1H]NMR as methylenedioxycinnamic acid (MDCA). This compound
 provided severe inhibition of curly cress (Lepidium sativum
 L.) root and shoot growth at concentrations of 25 ppm or
 above.
 
 
 169                                NAL Call. No.: QD415.A1J6
 Isolation and identification of allelochemicals that attract
 the larval parasitoid, Cotesia marginiventris (Cresson), to
 the microhabitat of one of its hosts.
 Turlings, T.C.J.; Tumlinson, J.H.; Heath, R.R.; Proveaux,
 A.T.; Doolittle, R.E.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2235-2259; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Cotesia marginiventris; Semiochemicals; Zea mays;
 Seedlings; Volatile compounds; Host-seeking behavior; Insect
 control; Biological control
 
 Abstract:  Volatiles released from corn seedlings on which
 beet armyworm larvae were feeding were attractive to females
 of the parasitoid, Cotesia marginiventris (Cresson), in flight
 tunnel bioassays. Analyses of the collected volatiles revealed
 the consistent presence of 11 compounds in significant
 amounts. They were: (Z)-3-hexenal, (E)-2-hexenal, (Z)-3-
 hexen-1-ol, (Z)-3-hexen-1-yl acetate, linalool,
 (3E)-4,8-dimethyl-1,3,7-nonatriene, indole, alpha-trans-
 bergamotene, (E)-beta-farnesene, (E)-nerolidol, and
 (3E,7E)-4,8,12-trimethyl-1,3,7,11-tridecatetraene. A synthetic
 blend of all 11 compounds was slightly less attractive to
 parasitoid females than an equivalent natural blend. However,
 preflight experience with the synthetic blend instead of
 experience with a regular plant-host complex significantly
 improved the response to the synthetic blend. Our results
 suggest that C marginiventris females, in their search for
 hosts, use a blend of airborne semiochemicals emitted by
 plants on which their hosts feed. The response to a particular
 odor blend dramatically increases after a parasitoid
 experiences it in association with contacting host by-
 products.
 
 
 170                                  NAL Call. No.: 450 P692
 Isolation and identification of lepidimoide, a new
 allelopathic substance from mucilage of germinated cress
 seeds.
 Hasegawa, K.; Mizutani, J.; Kosemura, S.; Yamamura, S.
 Rockville, MD : American Society of Plant Physiologists, 1926-
 ; 1992 Oct. Plant physiology v. 100 (2): p. 1059-1061; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Lepidium sativum; Seeds; Allelopathins;
 Mucilages; Seed germination; Allelopathy; Hypocotyls; Growth;
 Amaranthus caudatus
 
 Abstract:  A new allelopathic substance that promoted the
 shoot growth of different plant species but inhibited the root
 growth was isolated as an amorphous powder from mucilage of
 germinated cress (Lepidium sativum L.) seeds. This substance
 was identified as sodium
 2-O-rhamnopyranosyl-4-deoxy-threo-hex-4-enopyranosiduronate
 (designated lepidimoide) from the mass and the nuclear
 magnetic resonance and infrared spectra coupled with some
 chemical evidence. Lepidimoide promoted the hypocotyl growth
 of etiolated Amaranthus caudatus L. at concentrations higher
 than 3 micromolars and inhibited the root growth at
 concentrations higher than 100 micromolars. The growth-
 promoting activity in hypocotyls was 20 or 30 times as much as
 that of gibberellic acid.
 
 
 171                                 NAL Call. No.: 450 P5622
 Isolation of steroidal glycoalkaloids from Solanum incanum by
 two countercurrent chromatographic methods.
 Fukuhara, K.; Kubo, I.
 Oxford : Pergamon Press; 1991.
 Phytochemistry v. 30 (2): p. 685-687; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Kenya; Solanum incanum; Fruit; Medicinal plants;
 Products; Chemical composition; Glycoalkaloids; Allelopathins;
 Growth inhibitors; Chromatography; Spectral analysis
 
 Abstract:  Using a bioassay for inhibition of plant growth and
 a combination of two countercurrent chromatographies: rotation
 locular countercurrent chromatography and droplet
 countercurrent chromatography, two biologically active
 glycosidal alkaloids, solasonine and solamargine were isolated
 from fresh ripe fruit of Solanum incanum. The combination of
 these chromatographic techniques has established an efficient
 isolation of polar phytochemicals of steroidal glycoalkaloids.
 
 
 172                                NAL Call. No.: QD415.A1J6
 Isolation of substance from sweet potato (Ipomoea batatas)
 periderm tissue that inhibits seed germination.
 Peterson, J.K.; Harrison, H.F. Jr
 New York, N.Y. : Plenum Press; 1991 May.
 Journal of chemical ecology v. 17 (5): p. 943-951; 1991 May. 
 Includes references.
 
 Language:  English
 
 Descriptors: Ipomoea batatas; Periderm; Plant composition;
 Allelopathy; Seed germination; Inhibition; Weed control
 
 Abstract:  Chromatographic procedures were used to isolate
 inhibitors of seed germination from sweet potato root periderm
 tissue. The inhibitory activity of all fractions was monitored
 using a proso millet seed germination bioassay. A single HPLC
 peak, representing approximately 1.2% of the periderm dry
 weight, accounted for most of the inhibitory activity. The
 active fraction was labile in methanolic solution. Further
 fractionation of this peak by HPLC methods was not successful.
 In vitro seed germination dose-response relationships were
 established for the peak. The various seed species exhibited
 an extremely wide range of sensitivity. The I50 values were
 0.16, 0.013 and 0.011 mg/ml for redroot pigweed, velvetleaf,
 and proso millet, respectively. Tall morning glory was not
 inhibited by any concentration tested.
 
 
 173                                NAL Call. No.: QD415.A1J6
 Isothiocyanates as alleopathic compounds from Rorippa indica
 Hiern. (Cruciferae) roots.
 Yamane, A.; Fujikura, J.; Ogawa, H.; Mizutani, J.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Nov.
 Journal of chemical ecology v. 18 (11): p. 1941-1954; 1992
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Rorippa; Roots; Exudates; Allelopathy; Plant
 composition; Isothiocyanates; Bioassays; Weed control
 
 Abstract:  The ethyl acetate extracts of Rorippa indica Hiem.
 contained hirsutin, arabin, camelinin, and three novel omega-
 methylsulfonylalkyl isothiocyanates (n = 8, 9, and 10). These
 compounds severely inhibited lettuce (Lactuca sativa)
 hypocotyl and root growth at 0.1 mM or above. The precursor
 glucosinolates of hirsutin, arabin, and camelinin were
 isolated. Presence of the three omega-
 methylsulfonylalkylglucosinolates, along with other
 glucosinolates in the roots were verified by the isolation and
 identification of their desulfoderivatives. Using the
 continuous root exudate trapping apparatus and GC-MS, hirsutin
 and the three omega-methylsulfonylalkyl isothiocyanates were
 detected in the root exudates of R. indica, suggesting that
 these isothiocyanates are the primary candidate of
 allelopathic compounds contributing to the aggressiveness of
 this cruciferous weed.
 
 
 174                                 NAL Call. No.: 421 EN895
 Larval-damaged plants: source of volatile synomones that guide
 the parasitoid Cotesia marginiventris to the micro-habitat of
 its hosts. Turlings, T.C.J.; Tumlinson, J.H.; Eller, F.J.;
 Lewis, W.J. Dordrecht : Kluwer Academic Publishers; 1991 Jan.
 Entomologia experimentalis et applicata v. 58 (1): p. 75-82.
 ill; 1991 Jan. Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Crop damage; Leaves; Seedlings;
 Spodoptera exigua; Larvae; Volatile compounds;
 Allelochemicals; Insect attractants; Cotesia marginiventris;
 Host parasite relationships; Tunnels
 
 
 175                                 NAL Call. No.: 64.8 C883
 Medicarpin delays alfalfa seed germination and seedling
 growth. Dornbos, D.L. Jr; Spencer, G.F.; Miller, R.W.
 Madison, Wis. : Crop Science Society of America; 1990 Jan.
 Crop science v. 30 (1): p. 162-166; 1990 Jan.  Includes
 references.
 
 Language:  English
 
 Descriptors: Medicago sativa; Plant extracts; Medicarpin;
 Phytotoxicity; Growth inhibitors; Germination inhibitors; Seed
 germination; Seedling emergence; Allelopathy; Competitive
 ability; Bioassays; Abutilon theophrasti
 
 Abstract:  The objective of this study was to identify a
 compound(s) that may be responsible, at least in part, for
 alfalfa autotoxicity. We found that medicarpin (3-hydroxy-9-
 methoxypterocarpan) is a compound produced by alfalfa
 (Medicago sativa L.) that contributes to this. A clay loam
 soil from a declining alfalfa stand with two plants per 930
 cm2 and its 95% ethanol extract inhibited alfalfa emergence
 and seedling growth whereas extracted soil did not. The soil
 extracts contained at least 10 mg kg-1 medicarpin. Meicarpin,
 4-methoxy-medicarpin, sativan, and 5'-methoxysativan were
 isolated from alfalfa foliage with 95% ethanol, purified,
 identified by gas chromatography-mass spectroscopy and then
 exogenously applied to alfalfa and velvetleaf (Abutilon
 theophrasti L.) seeds using an agar bioassay. Medicarpin at 1
 X 10(-7) mole/seed reduced both alfalfa and velvetleaf
 seedling length by 39% after 72 h. Sativan and both methoxy
 derivatives had no affect on germination or seedling growth.
 In a timecourse study, 2 X 10(-7) mole/seed medicarpin delayed
 germination 12 h and seedling growth 44 h, during this time
 the alfalfa seedlings metabolized 85% of the medicarpin.
 Seedling growth then resumed at the control rate. Medicarpin
 is produced by mature alfalfa, is present in soil from a
 declining alfalfa stand, and is absorbed by alfalfa seedlings
 where it is metabolized and exerts a transient phytotoxic
 effect. Medicarpin may permit established alfalfa plants to
 control the ecology in their immediate proximity by inhibiting
 seedling establishment of nearby plants, thereby gaining a
 competitive advantage.
 
 
 176                                 NAL Call. No.: 421 EN895
 Metabolism and elimination of ingested allelochemicals in a
 holometabolous and a hemimetabolous insect.
 Smirle, M.J.; Isman, M.B.
 Dordrecht : Kluwer Academic Publishers; 1992 Feb.
 Entomologia experimentalis et applicata v. 62 (2): p. 183-190;
 1992 Feb. Includes references.
 
 Language:  English
 
 Descriptors: Melanoplus sanguinipes; Peridroma saucia;
 Allelochemicals; Excretion; Feces; Metabolic detoxification;
 Toxins
 
 
 177                                  NAL Call. No.: 450 P692
 Metabolism of L-canavanine and L-canaline in leguminous
 plants. Rosenthal, G.A.
 Rockville, Md. : American Society of Plant Physiologists; 1990
 Sep. Plant physiology v. 94 (1): p. 1-3; 1990 Sep.  Literature
 review.  Includes references.
 
 Language:  English
 
 Descriptors: Leguminosae; Metabolism; Canavanine; Arginine;
 Disease resistance; Plant competition; Biosynthesis;
 Metabolites; Literature reviews; Allelopathins
 
 Abstract:  Massive accumulation of L-canavanine, the 2-
 amino-4-(guanidinooxy) butyric acid structural analog of L-
 arginine, occurs in many legumes. Accumulation of large
 amounts of this nonprotein amino acid results in large part
 from canavanine's protective efficacy; it forms an effective
 chemical barrier to predation, disease, and even competition
 with other plants. Diversion of metabolic resources for the
 synthesis and storage of appreciable canavanine does not place
 an inordinate burden on the plant. Catabolism of this
 nonprotein amino acid provides respiratory carbon, generates
 essential primary metabolites, and ammoniacal nitrogen for the
 developing plant.
 
 
 178                                   NAL Call. No.: S601.D4
 Microbial degradation of plant materials and allelochemicals
 formation in different soils.
 Weyman-Kaczmarkowa, W.; Wojcik-Wojtkowiak, D.
 Amsterdam : Elsevier Scientific Publishing Company; 1992.
 Developments in agricultural and managed-forest ecology (25):
 p. 127-136; 1992.  In the series analytic: Humus its structure
 and role in agriculture and environment / edited by J. Kubat.
 Proceedings of the 10th Symposium Humus et Planta, August
 19-23, 1991, Prague, Czechoslovakia.  Literature review.
 Includes references.
 
 Language:  English
 
 Descriptors: Plant residues; Microbial degradation; Soil
 bacteria; Soil fungi; Allelochemicals; Allelopathy; Literature
 reviews
 
 Abstract:  The paper reviews current research concerning the
 accumulation of allelochemicals as influenced by microflora,
 kind and maturity of decomposed plant tissues as well as type
 of soil. On the basis of results from our investigations, the
 paper also discusses the interrelationship between the
 dynamics of bacterial and fungal development and formation of
 inhibitors from young rye and wheat plants undergoing
 decomposition in light and heavy soils. The interrelationship
 was found particularly conspicuous during the first several
 days of sample incubation. Rye autoinhibitory potential was
 higher than that of wheat. It was found that, in particular,
 zymogenic, macrotrophic bacteria and fungi are responsible for
 the formation of allelopathic inhibitors.
 
 
 179                              NAL Call. No.: QH541.5.D4J6
 Mineral nutrient content and turnover rate of Mesembryanthemum
 crystallinum in the north-western desert of Egypt.
 El-Darier, S.M.
 London : Academic Press; 1992 Apr.
 Journal of arid environments v. 22 (3): p. 219-230; 1992 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Egypt; Mesembryanthemum crystallinum; Mineral
 content; Nutrient content; Sodium; Calcium; Potassium;
 Magnesium; Spatial distribution; Vegetative period;
 Developmental stages; Litter (plant); Cycling; Allelopathy;
 Deserts; Leaves; Stems; Roots; Plant organs; Seasonal
 variation; Nutrient uptake
 
 
 180                                  NAL Call. No.: QL750.O3
 Mycorrhizal fungi and the nutrient ecology of three oldfield
 annual plant species.
 Koide, R.T.; Li, M.
 Berlin, W. Ger. : Springer International; 1991.
 Oecologia v. 85 (3): p. 403-412; 1991.  Includes references.
 
 Language:  English
 
 Descriptors: Abutilon theophrasti; Ambrosia artemisiifolia;
 Setaria (gramineae); Glomus etunicatum; Phosphorus; Roots;
 Allelopathy; Mycorrhizas
 
 
 181                                   NAL Call. No.: QD1.A45
 Natural phytotoxins as herbicides.
 Duke, S.O.; Lydon, J.
 Washington, D.C. : The Society; 1993.
 ACS Symposium series - American Chemical Society (524): p.
 110-124; 1993.  In the series analytic: Pest control with
 enhanced environmental safety / edited by S.O. Duke, J.J.
 Menn, and J.R. Plimmer.  Includes references.
 
 Language:  English
 
 Descriptors: Phytotoxins; Plant composition; Herbicidal
 properties; Allelopathy
 
 Abstract:  Natural products of plants and microbes offer a
 vast array of secondary compounds with biological activity,
 including phytotoxicity. Many of these compounds have the
 potential to be used directly as herbicides or as structural
 leads for new synthetic herbicides. Although natural compounds
 have made a large impact in the insecticide area, relatively
 few successes have been obtained with these compounds as
 herbicides. The most notable success is that of glufosinate.
 Use of natural products in a herbicide discovery strategy has
 been hindered by several problems. The number of options that
 must be considered in discovery and development of a natural
 product as a herbicide is larger than for a synthetic
 herbicide. Furthermore, the molecular complexity, limited
 environmental stability, and low herbicidal activity of many
 phytotoxic natural products are discouraging. Rediscovery of
 known natural phytotoxins can be time-consuming and expensive.
 However, advances in chemistry and biotechnology are
 increasing the speed and case with which humankind can
 discover and develop natural products as herbicides, while
 diminishing returns are being experienced with conventional
 herbicide discovery efforts based on "synthesize and screen"
 strategies.
 
 
 182                                   NAL Call. No.: 385 T29
 Novel sesquiterpene from bioactive fractions of cultivar
 sunflowers. Macias, F.A.; Varela, R.M.; Torres, A.; Molinillo,
 J.M.G. Oxford : Pergamon Press; 1993 Mar19.
 Tetrahedron letters v. 34 (12): p. 1999-2002; 1993 Mar19. 
 Includes references.
 
 Language:  English
 
 Descriptors: Helianthus annuus; Leaves; Allelochemicals; Plant
 composition; Sesquiterpenes; Isolation; Structure; Spectral
 data; X ray diffraction
 
 Abstract:  From the medium polar active fractions, we have
 isolated a sesquiterpene heliannuol A. It contains a
 previously unknown skeleton, heliannuol, whose structural
 elucidation was made based on spectroscopic technique and X-
 Ray diffraction analysis.
 
 
 183                                 NAL Call. No.: 64.8 C883
 Nutritional stresses and varietal resistance in rice: effects
 on whitebacked planthopper.
 Salim, M.; Saxena, R.C.
 Madison, Wis. : Crop Science Society of America; 1991 May.
 Crop science v. 31 (3): p. 797-805; 1991 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Oryza sativa; Pest resistance; Sogatella
 furcifera; Cultivars; Varietal susceptibility; Stress
 response; Nutrient deficiencies; Nitrogen; Phosphorus;
 Potassium; Nutrient content; Plant composition;
 Allelochemicals; Fecundity; Host parasite relationships
 
 Abstract:  Nutritional disorders can affect plant growth and a
 plant's susceptibility to pests. Our objective was to evaluate
 the effects of N, P, and K stresses on resistance of
 'IR2035-117-3' (IR2035) and susceptibility of 'Taichung Native
 1' (TN1) rice (Oryza sativa L.) plants to whitebacked
 planthopper, Sogatella furcifera (Horvath), when grown in
 nutrient solution in a phytotron at 29/21 degrees C
 (day/night), minimum 70% relative humidity, and natural
 daylight of 12 h. Nitrogen, P, or K stresses altered the
 chemical composition of rice plants. Deficiency of N, P, or K
 significantly reduced growth of rice plants. Allelochemical
 production decreased at low K (3 mg/kg) concentration. Insect
 food intake and assimilation growth, adult longevity,
 fecundity, and population increased significantly as N
 increased. In contrast, increases in K application adversely
 affected the biology and behavior of S. furcifera. Insect
 response to P-stressed plants was not consistent. Regardless
 of the levels of N, P, or K, the difference between the
 resistance of IR2035 and susceptibility of TN1 remained
 distinct. Mortality of first-instar nymphs was high on TN1
 plants treated with steam-distillate extracts of K-stressed or
 unstressed IR2035 plants when compared with plants treated
 with acetone or TN1 extract. Resistance to S. furcifera in
 rice cultivars thus may be enhanced by applying moderate rates
 of N and high doses of K.
 
 
 184                                NAL Call. No.: QD415.A1J6
 Ovipositional response of three Heliothis species
 (Lepidoptera: Noctuidae) to allelochemicals from cultivated
 and wild host plants.
 Mitchell, E.R.; Tingle, F.C.; Heath, R.R.
 New York, N.Y. : Plenum Press; 1990 Jun.
 Journal of chemical ecology v. 16 (6): p. 1817-1827; 1990 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Heliothis virescens; Heliothis subflexa;
 Helicoverpa zea; Allelochemicals; Oviposition deterrents;
 Plant extracts; Oviposition attractants; Gossypium; Nicotiana;
 Physalis; Desmodium; Insect control; Biological control
 
 Abstract:  The role of plant allelochemicals on the
 oviposition behavior of Heliothis virescens (F.), H. subflexa
 (Guenee), and H. zea (Boddie) was investigated in the
 laboratory using a "choice" bioassay system. Fresh young
 leaves of tobacco, Desmodium tortuosum (Swartz) de Candolle,
 groundcherry (Physalis angulata L.), and cotton (Gossypium
 hirsutum L.) squares (flower buds) were washed in methylene
 chloride or methanol, concentrated to 1 g equivalent of washed
 material, and applied to a cloth oviposition substrate. Each
 of the extracts-including groundcherry, a nonhost-stimulated
 oviposition by H. virescens. H. subflexa were stimulated to
 oviposit by groundcherry extract, its normal host, and extract
 from cotton squares, a nonhost. None of the extracts
 stimulated oviposition by H. zea, although all except
 groundcherry were from reported hosts. The sensitivity of the
 bioassay was confirmed by giving H. virescens and H. subflexa
 an opportunity to choose between extracts that showed
 stimulant qualities when tested independently versus only
 solvent-treated controls. In these tests, tobacco showed the
 highest level of stimulant activity for H. virescens;
 groundcherry exhibited the highest level of stimulation for H.
 subflexa.
 
 
 185                                  NAL Call. No.: 421 J825
 Oxidases in the gut of an aphid, Macrosiphum rosae (L.) and
 their relation to dietary phenolics.
 Peng, Z.; Miles, P.W.
 Exeter : Pergamon Press; 1991.
 Journal of insect physiology v. 37 (10): p. 779-787. ill;
 1991.  Includes references.
 
 Language:  English
 
 Descriptors: Rosa; Allelochemicals; Antifeedants; Catechin;
 Detoxification; Ingestion; Macrosiphum rosae; Digestive tract;
 Catechol oxidase; Honeydew; Peroxidase; Phenols; Saliva
 
 Abstract:  Catechol oxidase (EC 1.10.3.1) and peroxidase (EC
 1.11.1.7) were detected in the gut of the rose aphid. Both
 enzymes also occur in the saliva and catalyse the oxidation of
 catechin, a feeding deterrent that occurs in the parenchymal
 and vascular tissues of the rose. The oxidation products of
 catechin are phagostimulant, however, and the aphids will feed
 on tissues and on aqueous diets containing low concentrations
 of catechin. No catechin was detected in the gut, haemolymph
 or honeydew of aphids collected from roses but the presence in
 the gut and honeydew of phenolics that differed from those in
 the haemolymph was consistent with the ingestion and
 intraintestinal conversion of phenolics of plant origin.
 Evidence of the presence of catechin in the phloem sap was
 obtained, but whether the insects ingested phenols exclusively
 from phloem sap or from other tissues as well remained
 uncertain.
 
 
 186                                 NAL Call. No.: SD112.F67
 Partial suppression of pampas grass by other species at the
 early seedling stage.
 Gadgil, R.L.; Sandberg, A.M.; Allen, P.J.; Gallagher, S.S.
 Rotorua : The Institute; 1990.
 FRI bulletin - Forest Research Institute, New Zealand Forest
 Service (155): p. 120-127; 1990.  Paper presented at the
 "Conference on Alternatives to the Chemical Control of Weeds,"
 held July 25-27, 1989, Rotorua, New Zealand. Includes
 references.
 
 Language:  English
 
 Descriptors: New Zealand; Cortaderia selloana; Seedlings;
 Biological control; Allelopathy; Plant competition; Weed
 competition
 
 
 187                                NAL Call. No.: QD415.A1J6
 Phenolic acid content of soils from wheat-no till, wheat-
 conventional till, and fallow-conventional till soybean
 cropping systems.
 Blum, U.; Wentworth, T.R.; Klein, K.; Worsham, A.D.; King,
 L.D.; Gerig, T.M.; Lyu, S.W.
 New York, N.Y. : Plenum Press; 1991 Jun.
 Journal of chemical ecology v. 17 (6): p. 1045-1068; 1991 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Triticum aestivum; Glycine max; Phenolic acids;
 Soil chemistry; Allelopathy; Fallow systems; Tillage; Weed
 control; Biological control
 
 Abstract:  Soil core (0-2.5 and/or 0.10 cm) samples were taken
 from wheat-no till, wheat-conventional till, and fallow-
 conventional till soybean cropping systems from July to
 October of 1989 and extracted with water in an autoclave. The
 soil extracts were analyzed for seven common phenolic acids
 (p-coumaric, vanillic, p-hydroxybenzoic, syringic, caffeic,
 ferulic, and sinapic; in order of importance) by high-
 performance liquid chromatography. The highest concentration
 observed was 4 micrograms/g soil for p-coumaric acid. Folin &
 Ciocalteu's phenol reagent was used to determine total
 phenolic acid content. Total phenolic acid content of 0- to
 2.5-cm core samples was approximately 34% higher than that of
 the 0- to 10-cm core samples. Phenolic acid content of 0-to
 2.5-cm core samples from wheat-no till systems was
 significantly higher than those from all other cropping
 systems. Individual phenolic acids and total phenolic acid
 content of soils were highly correlated. The last two
 observations were confirmed by principal component analysis.
 The concentrations were confirmed by principal component
 analysis of individual phenolic acids extracted from soil
 samples were related to soil pH, water content of soil
 samples, total soil carbon, and total soil nitrogen. Indirect
 evidence suggested that phenolic acids recovered by the water-
 autoclave procedure used came primarily from bound forms in
 the soil samples.
 
 
 188                                NAL Call. No.: QD415.A1J6
 Phenylacetic acid as a phytotoxic compound of corn pollen.
 Anaya, A.L.; Hernandez-Bautista, B.E.; Jimenez-Estrada, M.;
 Velasco-Ibarra, L. New York, N.Y. : Plenum Publishing
 Corporation; 1992 Jun. Journal of chemical ecology v. 18 (6):
 p. 897-905; 1992 Jun.  Includes references.
 
 Language:  English
 
 Descriptors: Zea mays; Zea mexicana; Pollen; Chemical
 composition; Allelopathy; Phytotoxicity; Amaranthus
 leucocarpus; Echinochloa crus-galli
 
 Abstract:  Phenylacetic acid (PAA), one of the phytotoxic
 compounds in corn (Zea mays) pollen, was identified by GC-MS
 and by direct comparison with a pure commercial sample of PAA.
 Bioassays were carried out by testing whole pollen, methylene
 chloride extract of the pollen, and pure PAA on germination
 and radical growth of Amaranthus leucocarpus and Echinochloa
 crusgalli. The effect of corn pollen was compared with that of
 Zea mexicana (Teosinte), one of the wild relatives of
 cultivated maize.
 
 
 189                                   NAL Call. No.: 26 T756
 Phytotoxic effects of tree crops on germination and radicle
 extension of some food crops.
 Bhatt, B.P.; Chauhan, D.S.; Todaria, N.P.
 London : Whurr Publishers Ltd; 1993.
 Tropical science v. 33 (1): p. 69-73; 1993.  Includes
 references.
 
 Language:  English
 
 Descriptors: Uttar pradesh; Forest trees; Bark; Leaves;
 Leachates; Plant extracts; Agroforestry; Allelopathy; Food
 crops; Phytotoxicity; Radicles; Seed germination; Glycine max;
 Macrotyloma uniflorum; Phaseolus lunatus; Vigna mungo
 
 
 190                                NAL Call. No.: QD415.A1J6
 Phytotoxicity of sorgoleone found in grain sorghum root
 exudates. Einhellig, F.A.; Souza, I.F.
 New York, N.Y. : Plenum Press; 1992 Jan.
 Journal of chemical ecology v. 18 (1): p. 1-11; 1992 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Sorghum; Root exudates; Phytotoxins; Allelopathy
 
 Abstract:  Root exudates of Sorghum bicolor consist primarily
 of a dihydroquinone that is quickly oxidized to a p-
 benzoquinone named sorgoleone. The aim of this investigation
 was to determine the potential activity of sorgoleone as an
 inhibitor of weed growth. Bioassays showed 125 micromolar
 sorgoleone reduced radicle elongation of Eragrostis tef. In
 liquid culture, 50-micromolar sorgoleone treatments stunted
 the growth of Lemna minor. Over a 10-day treatment period, 10
 micromolar sorgoleone in the nutrient medium reduced the
 growth of all weed seedlings tested: Abutilon theophrasti,
 Datura stramonium, Amaranthus retroflexus, Setaria viridis,
 Digitaria sanguinalis, and Echinochloa crusgalli. These data
 show sorgoleone has biological activity at extremely low
 concentrations, suggesting a strong contribution to Sorghum
 allelopathy.
 
 
 191                                NAL Call. No.: QD415.A1J6
 Plant growth regulatory activities of artemisinin and its
 related compounds. Chen, P.K.; Leather, G.R.
 New York, N.Y. : Plenum Press; 1990 Jun.
 Journal of chemical ecology v. 16 (6): p. 1867-1876; 1990 Jun. 
 Includes references.
 
 Language:  English
 
 Descriptors: Artemisia annua; Allelopathy; Sesquiterpenoid
 lactones; Seed germination; Phytotoxicity; Bioassays
 
 Abstract:  Artemisinin, a sesquiterpene lactone produced in
 the leaves of Artemisia annua, was evaluated for its
 phytotoxicity in mono- and dicotyledonous plants. Artemisinin
 inhibited seed germination, seedling growth, and root
 induction in all species tested. The concentration of
 artemisinin required for 50% inhibition of Lemna minor growth
 was 5 micromolar. Inhibitory plant responses appeared to
 require the endoperoxide moiety of this compound since similar
 chemicals without endoperoxide, deoxyartemisinin, arteannuic
 acid, and arteannuin B, were less phytotoxic. In L. minor,
 artemisinin and arteannuic acid caused the leakage of proteins
 into the growth medium, suggesting the site of activity was at
 the plant cell membrane.
 
 
 192                                NAL Call. No.: QD415.A1J6
 Plant structures of manipulating predator-prey interactions
 through allelochemicals: prospects for application in pest
 control. Dicke, M.; Sabelis, M.W.; Takabayashi, J.; Bruin, J.;
 Posthumus, M.A. New York, N.Y. : Plenum Press; 1990 Nov.
 Journal of chemical ecology v. 16 (11): p. 3091-3118; 1990
 Nov.  Proceedings of an International Symposium:
 Semiochemicals and Pest Control--Prospects for New
 Applications, October 16-19, 1989, Wageningen, The
 Netherlands.  Includes references.
 
 Language:  English
 
 Descriptors: Plant composition; Host plants; Pest resistance;
 Allelochemicals; Predatory arthropods; Searching behavior;
 Predator prey relationships; Defense; Insect control;
 Biological control
 
 Abstract:  To understand the role of allelochemicals in
 predator-prey interactions it is not sufficient to study, the
 behavioral responses of predator and prey. One should
 elucidate the origin of the allelochemicals and be aware that
 it may be located at another trophic level. These aspects are
 reviewed for predator-prey interactions in general and
 illustrated in detail for interactions between predatory mites
 and herbivorous mites. In the latter system there is
 behavioral and chemical evidence for the involvement of the
 host plant in production of volatile allelochemicals upon
 damage by the herbivores with the consequence of attracting
 predators. These volatiles not only influence predator
 behavior, but also prey behavior and even the attractiveness
 of nearby plants to predators. Herbivorous mites disperse away
 from places with high concentrations of the volatiles, and
 undamaged plants attract more predators when previously
 exposed to volatiles from infested conspecific plants a rather
 than from uninfested plants. The latter phenomenon may well be
 an example of plant-to-plant communication. The involvement of
 the host plant is probably not unique to the predator-
 herbivore-plant system under study. it may well be widespread
 since it makes sense from an evolutionary point of view. If
 so, prospects for application in pest control are wide open.
 These are discussed, and it is concluded that crop protection
 in the future should include tactics whereby man becomes an
 ally to plants in their strategies to manipulate predator-prey
 interactions through allelochemicals.
 
 
 193                                   NAL Call. No.: QD1.A45
 Plant-allelochemical-adapted glutathione transferases in
 Lepidoptera. Yu, S.J.
 Washington, D.C. : The Society; 1992.
 ACS Symposium series - American Chemical Society (505): p.
 174-190; 1992.  In the series analytic: Molecular mechanisms
 of insecticide resistance / edited by C.A. Mullin and J.G.
 Scott.  Includes references.
 
 Language:  English
 
 Descriptors: Spodoptera frugiperda; Trichoplusia ni;
 Anticarsia gemmatalis; Helicoverpa zea; Heliothis virescens;
 Plant composition; Allelochemicals; Metabolic detoxification;
 Glutathione transferase; Insect control
 
 Abstract:  Glutathione transferases metabolized toxic
 allelochemicals, including alpha, beta-unsaturated carbonyl
 compounds, isothiocyanates and organothiocyanates in
 lepidopterous insects. These transferase activities in the
 specialist velvetbean caterpillar are lower than in the
 generalist fall armyworm; the activity toward the
 isothiocyanates in the crucifer-adapted cabbage looper was 2-
 to 6-fold higher than that in the fall armyworm. Host plants
 such as crucifers and umbellifers, and allelochemicals such as
 coumarins, indoles, flavonoids, isothiocyanates and
 monoterpenes induced glutathione transferases in these
 insects. The highly polyphagous Lepidoptera, fall armyworm and
 corn earworm, possessed multiple glutathione transferases
 containing six and four isozymes, respectively, whereas the
 more specialized Lepidoptera, tobacco budworm, cabbage looper
 and velvetbean caterpillar, had a single form of the enzyme.
 The results suggest that glutathione transferases play an
 important role in allelochemical resistance in phytophagous
 Lepidoptera.
 
 
 194                                NAL Call. No.: SB950.A1P3
 Plants with insecticidal activities against four major insect
 pests in Pakistan.
 Anwar, T.; Jabbar, A.; Khalique, F.; Tahir, S.; Shakeel, M.A.
 London : Taylor & Francis; 1992 Oct.
 Tropical pest management v. 38 (4): p. 431-437; 1992 Oct. 
 Includes references.
 
 Language:  English
 
 Descriptors: Pakistan; Insect pests; Mortality; Plant pests;
 Allelochemicals; Insect control; Insecticidal action;
 Insecticidal plants; Plant extracts
 
 
 195                                NAL Call. No.: QD415.A1J6
 Possible ecological significance of within-fruit and seed
 furocoumarin distribution in two Psoralea species.
 Cappelletti, E.M.; Innocenti, G.; Caporale, G.
 New York, N.Y. : Plenum Press; 1992 Feb.
 Journal of chemical ecology v. 18 (2): p. 155-164; 1992 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Psoralea macrostachya; Psoralea; Fruits; Seeds;
 Chemical composition; Allelopathy; Coumarins; Seed
 germination; Inhibition; Defense
 
 Abstract:  Furocoumarin distribution patterns in the fruits
 and seeds of Psoralea macrostachya and P. onobrychis were
 investigated. Both species contain the linear furocoumarin
 psoralen and its angular isomer, angelicin. In the
 monospermous indehiscent fruit of P. macrostachya,
 furocoumarins occur in the pericarp and all seed parts. In P.
 onobrychis, the pericarp of which is easily detached at
 ripeness, no furocoumarins were found in the pericarp tissues
 and only traces occur in the embryo axis; cotyledons are the
 preferential accumulation site. The within-fruit and -seed
 furocoumarin variations associated with the developmental
 stages of fruit were followed in P. onobrychis, in view of
 changes in the defensive value of the pericarp before and
 after ripening. Rapid furocoumarin biosynthesis after
 fertilization was observed in both pericarp and seed; ripening
 is associated with furocoumarin decrease in the seed and
 complete disappearance in the pericarp tissues. Such findings
 are consistent with the chemical defense role of these
 substances. The cooccurrence of linear and angular isomers
 seems to be a chemical marker of the genus Psoralea: the
 biosynthetic pathway leading to the angular isomer as an
 evolutionary response to selective pressure from herbivore
 insects is suggested.
 
 
 196                                NAL Call. No.: QD415.A1J6
 Potential allelochemicals from Pistia stratiotes L.
 Aliotta, G.; Monaco, P.; Pinto, G.; Pollio, A.; Previtera, L.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2223-2234; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Pistia stratiotes; Plant composition;
 Allelochemicals; Growth inhibitors; Algae
 
 Abstract:  Among the substances isolated from ethyl ether
 extract of Pistia stratiotes L., linoleic acid, gamma-
 linolenic acid,
 (12R,9Z,13E,15Z)-12-hydroxy-9,13,15-octadecatrienoic acid,
 (9S,10E,12Z,15Z)-9-hydroxy-10,12,15-octadecatrienoic acid,
 alpha-asarone, and 24S-ethyl-4,22-cholestadiene-3,6-dione were
 found to inhibit the growth of some microalgae in solid
 medium. Toxicity of alpha-asarone on the growth of sensitive
 algal strains in liquid medium is discussed.
 
 
 197                                 NAL Call. No.: 450 P5622
 Potential allelopathic activity of several sesquiterpene
 lactone models. Macias, F.A.; Galindo, J.C.G.; Massanet, G.M.
 Oxford : Pergamon Press; 1992 Jun.
 Phytochemistry v. 31 (6): p. 1969-1977; 1992 Jun.  Includes
 references.
 
 Language:  English
 
 Descriptors: Lactuca sativa; Bioassays; Sesquiterpenoid
 lactones; Allelopathins
 
 Abstract:  A collection of 12 natural and synthetic
 sesquiterpene lactones with eudesmanolide, melampolide,
 cis,cis-germacranolide, and guaianolide skeletons have been
 prepared and tested as allelochemicals. The effect of a series
 of aqueous solutions at 10(-4)-10(-9) M of this collection is
 evaluated. The specific structural requirements related to
 their activity is discussed. The natural sesquiterpene
 lactones soulangianolide A, melampomagnolide A and B,
 zaluzanin C and isozaluzanin C have been synthesized from
 costunolide, parthenolide and dehydrocostuslactone using SeO2,
 and tert-butylhydroperoxide. The structures of the synthetic
 compounds were established by NMR spectroscopy.
 
 
 198                                  NAL Call. No.: SB599.C8
 Potential allelopathic influence of certain pasture weeds.
 Smith, A.E.
 Guildford : Butterworths; 1990 Dec.
 Crop protection v. 9 (6): p. 410-414; 1990 Dec.  Includes
 references.
 
 Language:  English
 
 Descriptors: Medicago sativa; Lolium multiflorum; Crop weed
 competition; Pastureplants; Weeds; Allelopathy; Eupatorium
 capillifolium; Anthemis cotula; Seedlings; Plant development;
 Bioassays
 
 
 199                                 NAL Call. No.: 79.9 C122
 Potential for weed control with allelopathy in turfgrass.
 Elmore, C.L.
 Fremont, Calif. : California Weed Conference; 1990.
 Proceedings - California Weed Conference (42): p. 214-216;
 1990.  Meeting held January 15-17, 1990, San Jose, California.
 
 Language:  English
 
 Descriptors: Lawns and turf; Allelopathy; Weed control;
 Biological control
 
 
 200                                NAL Call. No.: QD415.A1J6
 Prospects of antifeedant approach to pest control--a critical
 review. Jermy, T.
 New York, N.Y. : Plenum Press; 1990 Nov.
 Journal of chemical ecology v. 16 (11): p. 3151-3166; 1990
 Nov.  Proceedings of an International Symposium:
 Semiochemicals and Pest Control--Prospects for New
 Applications, October 16-19, 1989, Wageningen, The
 Netherlands.  Includes references.
 
 Language:  English
 
 Descriptors: Antifeedants; Chemoreceptors; Allelochemicals;
 Electrophysiology; Integrated pest management; Screening;
 Field tests; Structure activity relationships; Toxicity;
 Persistence; Insect control; Biological control
 
 Abstract:  The increasing efforts to develop environmentally
 safer pest control methods have attracted the attention of
 many authors towards the use of antifeedants. This review is a
 critical survey of the most important publications issued
 during the 1980s, especially dealing with sensory
 physiological and behavioral studies, structure-activity
 aspects, screening methods, sources of antifeedants, and
 aspects of practical application. Important but neglected
 areas of research as well as future tasks are stressed.
 
 
 201                                NAL Call. No.: QD415.A1J6
 Quinolizidine alkaloids in Genista acanthoclada and its
 holoparasite, Cuscuta palaestina.
 Wink, M.; Witte, L.
 New York, N.Y. : Plenum Publishing Corporation; 1993 Mar.
 Journal of chemical ecology v. 19 (3): p. 441-448; 1993 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Genista; Cuscuta; Plant composition;
 Quinolizidine alkaloids; Phloem loading; Defense mechanisms;
 Allelochemicals
 
 Abstract:  About 20 quinolizidine alkaloids were identified in
 Genista acanthoclada by capillary GLC and GLC-MS, such as
 sparteine, 11,12-dehydrosparteine, retamine, N-methylcytisine,
 cytisine, 17-oxosparteine, lupanine, alpha-isolupanine, 5,6-
 dehydrolupanine, 10-oxosparteine, N-carbomethoxycytisine, 17-
 oxoretamine, N-formylcytisine, N-acetylcytisine, and
 anagyrine. Its phloem-feeding holoparasite Cuscuta palaestina
 contained alkaloids too, such as sparteine, 11,12-
 dehydrosparteine, retamine, N-methylcytisine, cytisine, 17-
 oxosparteine, lupanine, N-carbomethoxycytisine, and anagyrine.
 Whereas sparteine, retamine, 17-oxosparteine, and cytisine are
 the main alkaloids of G. acanthoclada, lupanine, cytisine, N-
 methylcytisine, and anagyrine are abundant and enriched in C.
 palaestina. Since these alkaloids figure as antiherbivoral
 chemical defense compounds in Genista, it is assumed that the
 parasite can exploit the acquired allelochemicals for its own
 protection.
 
 
 202                                 NAL Call. No.: QK900.J67
 Recruitment pattern of Rhus integrifolia populations in
 periods between fire in chaparral.
 Lloret, F.; Zedler, P.H.
 Knivsta, Sweden : Opulus Press; 1991 Apr.
 Journal of vegetation science v. 2 (2): p. 217-230; 1991 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: California; Rhus; Allelopathy; Chaparral; Coastal
 plant communities; Fire ecology; Plant succession; Seed banks;
 Seed dispersal; Seed predation
 
 
 203                                NAL Call. No.: QD415.A1J6
 Relationships between chemical structure and inhibitory
 activity of C6 through C9 volatiles emitted by plant residues.
 Bradow, J.M.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2193-2212; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Allium cepa; Daucus carota; Lycopersicon
 esculentum; Allelopathy; Plant composition; Volatile
 compounds; Seed germination
 
 Abstract:  Leaf, stem, flower, fruit, and root residues of a
 wide variety of plants have been reported to emit bioactive n-
 alkanes, 2-alkanols, n-alkanals, 2-alkenals, 2-alkanones, and
 n-alkanoic acids containing from six to nine carbon atoms.
 During a 72-hr exposure to the vapor phase of these compounds
 (6.9, 20.8 or 34.4 micromoles/liter), germination of onion,
 carrot, and tomato seeds was inhibited to varying degrees. The
 extent of inhibition caused by these plant residue volatiles
 depended upon the compound type and concentration, carbon-
 chain length, and the seed species tested. Tomato seeds
 recovered more fully from exposure to these volatile
 inhibitors than did those of carrot and onion. Degree of
 recovery in all three species depended on the type and
 concentration of volatile present. The C7 and C8 compounds
 were the most inhibitory among these 24 volatiles. Of the six
 classes of chemicals examined, the 2-alkenals were the most
 inhibitory, followed by the 2-alkanols, n-alkanals, and 2-
 alkanones, which were equally effective as seed germination
 inhibitors. The straight-chain alkanes and alkanoic acids were
 relatively noninhibitory. Tests of a C7 and C9 alkadienal
 indicated that the C7 compound was the more inhibitory.
 
 
 204                                  NAL Call. No.: 421 J822
 Relationships of glands, cotton square terpenoid aldehydes,
 and other allelochemicals to larval growth of Heliothis
 virescens (Lepidoptera: Noctuidae).
 Hedin, P.A.; Parrott, W.L.; Jenkins, J.N.
 Lanham, Md. : Entomological Society of America; 1992 Apr.
 Journal of economic entomology v. 85 (2): p. 359-364; 1992
 Apr.  Includes references.
 
 Language:  English
 
 Descriptors: Mississippi; Gossypium hirsutum; Cultivars;
 Lines; Pest resistance; Plant glands; Susceptibility;
 Terpenoids; Aldehydes; Allelochemicals; Heliothis virescens;
 Larvae; Growth
 
 Abstract:  Female moths of the tobacco budworm, Heliothis
 virescens (F.), oviposit in terminals of the cotton plant,
 Gossypium hirsutum (L.). The hatched larvae feed in the
 terminal area, then migrate to small squares (buds) where they
 feed and finally burrow into and feed on the anthers, where
 they grow rapidly. They attempt to avoid feeding on gossypol
 glands during the first 48 h after hatching. When tobacco
 budworm neonate larvae were fed squares of highly glanded
 lines, growth was decreased by 25-75 %. The number of glands
 in calyx and bract tissues of squares of resistant lines was
 significantly higher than in susceptible lines. The difference
 was greatest in the calyx crown where the ratio in resistant
 to susceptible lines was 10-20 fold. The calyx crown of highly
 glanded resistant lines also was high in terpenoid aldehydes.
 High pressure liquid chromatography data showed that the
 gossypol content of susceptible and resistant glanded lines is
 equal, whereas three other terpenoid aldehydes,
 hemigossypolone and heliocides H1 and H2 are greatly increased
 in resistant lines, and they are presumably more closely
 associated with resistance.
 
 
 205                                   NAL Call. No.: QK1.A28
 Relative effects of Prosopis juliflora swartz and Prosopis
 cineraria (L.) druce on seed germination and seedling growth.
 Goel, U.; Nathawat, G.S.
 Meerut, India : Society for Advancement of Botany; 1990 Jun.
 Acta botanica Indica v. 18 (1): p. 76-79; 1990 Jun.  Includes
 references.
 
 Language:  English
 
 Descriptors: India; Prosopis juliflora; Prosopis cineraria;
 Crotalaria medicaginea; Indigofera; Seed germination;
 Seedlings; Growth; Plant extracts; Growthpromoters; Growth
 retardants; Allelopathins; Plant interaction
 
 
 206                                NAL Call. No.: QD415.A1J6
 Release of allelochemical agents from litter, throughfall, and
 topsoil in plantations of Eucalyptus globulus labill in Spain.
 Molina, A.; Reigosa, M.J.; Carballeira, A.
 New York, N.Y. : Plenum Press; 1991 Jan.
 Journal of chemical ecology v. 17 (1): p. 147-160; 1991 Jan. 
 Includes references.
 
 Language:  English
 
 Descriptors: Spain; Eucalyptus globulus; Allelopathy; Lactuca
 sativa; Phytotoxicity; Leachates; Soil
 
 Abstract:  Natural leachates of Eucalyptus globulus
 (throughfall, stemflow, and soil percolates) were collected
 daily during rainy spells in the vegetative period (February-
 July), and their effects on the germination and radicle growth
 of Lactuca sativa were measured. Concurrently, the effects of
 L. sativa of topsoil and leachates from decaying litter were
 determined. The results suggest that toxic allelochemicals
 released by Eucalyptus globulus may influence the composition
 and structure of the understory of the plantation and that
 this effect is attributable mainly to the decomposition
 products of decaying litter rather than to aerial leachates.
 The soil may neutralize or dilute allelopathic agents, at
 least below the top few cms.
 
 
 207                                   NAL Call. No.: 410 M58
 Replacement of Cakile edentula by C. maritima in the strand
 habitat of California.
 Boyd, R.S.; Barbour, M.G.
 Notre Dame, Ind. : University of Notre Dame, 1909-; 1993 Oct.
 The American midland naturalist v. 130 (2): p. 209-228; 1993
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: California; Cabt; Cruciferae; Coastal plant
 communities; Allelopathy; Duneland plants; Ecosystems;
 Habitats; Herbivores; Invasion; Plant competition; Plant
 succession; Survival
 
 
 208                                   NAL Call. No.: QD1.A45
 Resistance to plant allelochemicals in Heliothis virescens
 (Fabricius). Rose, R.L.; Gould, F.; Levi, P.; Konno, T.;
 Hodgson, E.
 Washington, D.C. : The Society; 1992.
 ACS Symposium series - American Chemical Society (505): p.
 137-148; 1992.  In the series analytic: Molecular mechanisms
 of insecticide resistance / edited by C.A. Mullin and J.G.
 Scott.  Includes references.
 
 Language:  English
 
 Descriptors: Heliothis virescens; Allelochemicals; Resistance;
 Cytochrome p-450; Resistance mechanisms; Insect control
 
 Abstract:  Potential metabolic routes for the resistance of
 tobacco budworm (TBW) larvae to host plant allelochemicals
 nicotine, 2-tridecanone and quercetin were explored. Midgut
 preparations from larvae resistant to nicotine and 2-
 tridecanone had elevated levels of cytochrome P450 which were
 associated with significant increases in metabolism for five
 of six monooxygenase substrates. In quercetin tolerant larvae,
 metabolism of two monooxygenase substrates was significantly
 enhanced although no increase in P450 content was observed.
 Glutathione transferases and esterases did not appear to be
 involved in the resistance of any of the strains examined.
 Patterns of substrate oxidations varied between strains and
 inducing agents, suggesting that different isozymes of P450
 are associated with resistance and induction.
 
 
 209                                   NAL Call. No.: 450 M99
 Role of Acremonium endophyte of fescue on inhibition of
 colonization and reproduction of mycorrhizal fungi.
 Guo, B.Z.; Hendrix, J.W.; An, Z.Q.; Ferriss, R.S.
 Bronx, N.Y. : The New York Botanical Garden; 1992 Nov.
 Mycologia v. 84 (6): p. 882-885; 1992 Nov.  Includes
 references.
 
 Language:  English
 
 Descriptors: Festuca arundinacea; Acremonium coenophialum;
 Endophytes; Glomus mosseae; Glomus macrocarpum; Mycorrhizal
 fungi; Infectivity; Inhibition; Allelopathy; Allelopathins;
 Alkaloids; Toxic exudates
 
 
 210                                NAL Call. No.: QH540.E288
 The role of allelopathy in agroecosystems: studies from
 tropical Taiwan. Chou, C.H.
 New York, N.Y. : Springer-Verlag; 1990.
 Ecological studies : analysis and synthesis v. 78: p. 104-121;
 1990.  In the series analytic: Agroecology : Researching the
 Ecological Basis for Sustainable Agriculture / edited by
 Stephen R. Gliessman.  Includes references.
 
 Language:  English
 
 Descriptors: Taiwan; Allelopathy; Cropping systems;
 Environmental factors; Ecosystems; Tropics
 
 
 211                                NAL Call. No.: QD415.A1J6
 Role of avian trigeminal sensory system in detecting coniferyl
 benzoate, a plant allelochemical.
 Jakubas, W.J.; Mason, J.R.
 New York, N.Y. : Plenum Press; 1991 Nov.
 Journal of chemical ecology v. 17 (11): p. 2213-2221; 1991
 Nov.  Includes references.
 
 Language:  English
 
 Descriptors: Sturnus vulgaris; Antifeedants; Plant
 composition; Senses; Chemoreceptors; Pest control; Biological
 control
 
 Abstract:  Coniferyl benzoate, a secondary metabolite found in
 quaking aspen (Populus tremuloides) and other plants, is an
 avian feeding deterrent of ecological and potential commercial
 importance. This study was conducted to determine if coniferyl
 benzoate is a trigeminal stimulant for birds and to ascertain
 if trigeminal chemoreception of coniferyl benzoate can mediate
 avian feeding behavior. Five European starlings (Sturnus
 vulgaris) with bilateral nerve cuts ophthalmic branch of the
 trigeminal nerve) and four starlings that had sham surgeries
 were fed a commercial diet treated with coniferyl benzoate.
 Birds receiving bilateral nerve cuts ate significantly more
 feed than intact birds, indicating trigeminal detection of
 coniferyl benzoate and trigeminal mediation of feeding
 behavior. In the past, trigeminal chemoreception has not been
 recognized as important in the detection of plant secondary
 metabolites despite the irritant or astringent properties of a
 number of them.
 
 
 212                                  NAL Call. No.: QL495.A7
 Role of superoxide dismutase in the protection and tolerance
 to the prooxidant allelochemical quercetin in Papilio
 polyxenes, Spodoptera eridania, and Trichoplusia ni.
 Pritsos, C.A.; Pastore, J.; Pardini, R.S.
 New York, N.Y. : Wiley-Liss; 1991.
 Archives of insect biochemistry and physiology v. 16 (4): p.
 273-282; 1991. Includes references.
 
 Language:  English
 
 Descriptors: Papilio polyxenes; Spodoptera eridania;
 Trichoplusia ni; Larvae; Superoxide dismutase; Quercetin
 
 Abstract:  Larvae of the black swallowtail butterfly, Papilio
 polyxenes, the southern armyworm, Spodoptera eridania, and the
 cabbage looper, Trichoplusia ni, have different feeding habits
 and dietary breadth, which contributes to differences in their
 exposure and tolerance to dietary prooxidant allelochemicals.
 The antioxidant enzyme activities of larvae of these insects
 have been previously determined, with the levels being P
 polyxenes > S. eridania > T ni. The relative activities of
 these antioxidant enzymes are consistent with the relative
 exposure of these insects to prooxidants. This suggests that
 the antioxidant enzymes may play a role in the defense against
 allelochemical toxicity in these insects. Dietary
 diethlydithiocarbamate (DETC), a copper chelating agent and
 superoxide dismutase (SOD) inhibitor, was shown to inhibit SOD
 in all three insects. Toxicological studies were conducted
 using four diets for each insect. The standard diets for each
 insect were supplemented with either control (solvent),
 quercetin (a prooxidant), DETC, or DETC plus quercetin.
 Nontoxic doses of each compound for each insect were used.
 inhibition of SOD in P. polyxenes and S. eridania dramatically
 increased quercetin-induced toxicity as measured by relative
 growth and consumption rates in these species. DETC had no
 effect on quercetin toxicity in T ni. These results elucidate
 the important role of SOD in the prooxidant allelochemical
 defense of insects.
 
 
 213                                NAL Call. No.: QD415.A1J6
 Role of the isoflavonoid coumestrol in the constitutive
 antixenosic properties of "Davis" soybeans against an
 oligophagous insect, the Mexico bean beetle. Burden, B.J.;
 Norris, D.M.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Jul.
 Journal of chemical ecology v. 18 (7): p. 1069-1081; 1992 Jul. 
 Includes references.
 
 Language:  English
 
 Descriptors: Glycine max; Allelochemicals; Flavonoids; Feeding
 behavior; Epilachna varivestis; Insect control
 
 Abstract:  The antixenosic properties of the isoflavonoid,
 coumestrol, were tested in dual-choice leaf disk bioassays
 with the Mexican bean beetle (Epilachna varivestis Mulsant).
 E. varivestis preferred the methanol-treated (solvent control)
 disk when the coumestrol concentration was 1.8 or 0.9
 micrograms/leaf disk. No preference was observed between the
 coumestrol-treated and the solvent-control disks when the
 coumestrol concentration was higher, at 3.6, or lower, at 0.45
 micrograms/leaf disk. Coumestrol alone clearly is not
 responsible for the significant constitutive antixenosic
 properties of "Davis" soybeans, Glycine max (L.) Merrill,
 because the amount of coumestrol in these plants is
 significantly less than the minimum concentration which was
 antixenosic in this study. However, it might contribute to a
 constitutive antixenosis in "Davis" involving a profile of
 allelochemicals. A computer-aided densitometer, adapted to
 measure the leaf disk area, increased the resolution of the
 leaf area 250 (x)-fold as compared to the standard LI-COR leaf
 area meter.
 
 
 214                                 NAL Call. No.: 450 P5622
 Root exudates of wild oats: allelopathic effect on spring
 wheat. Perez, F.J.; Ormeno-Nunez, J.
 Oxford : Pergamon Press; 1991.
 Phytochemistry v. 30 (7): p. 2199-2202; 1991.  Includes
 references.
 
 Language:  English
 
 Descriptors: Avena fatua; Root exudates; Triticum aestivum;
 Allelopathins; Seedlings; Roots; Coleoptiles; Growth
 inhibitors; Herbicidal properties; Phytotoxicity
 
 Abstract:  Root exudates from the undisturbed root system of
 wild oats Avena fatua were collected by a modification of the
 Tang and Young method. Exudates inhibited root and coleoptile
 growth of spring wheat seedlings (Triticum aestivum).
 Scopoletin, coumarin, p-hydroxybenzoic and vanillic acid were
 tentatively identified from the root exudates by HPLC.
 
 
 215                                 NAL Call. No.: 450 B6527
 Root flavonoids.
 Rao, A.S.
 Bronx, N.Y. : New York Botanical Garden; 1990 Jan.
 The Botanical review v. 56 (1): 90 p.; 1990 Jan.  Includes
 references.
 
 Language:  English
 
 Descriptors: Chemical constituents of plants; Root analysis;
 Flavonoids; Nutrient uptake; Growth rate; Graviperception;
 Allelopathy; Nitrogen fixation; Symbiosis; Biosynthesis; Plant
 protection; Plant metabolism; Plant pigments; Medicinal plants
 
 
 216                                   NAL Call. No.: 4 AM34P
 Scanning electron microscopy for studying root morphology and
 anatomy in alfalfa autotoxicity.
 Hegde, R.S.; Miller, D.A.
 Madison, Wis. : American Society of Agronomy; 1992 Jul.
 Agronomy Journal v. 84 (4): p. 618-620; 1992 Jul.  Includes
 references.
 
 Language:  English
 
 Descriptors: Medicago sativa; Allelopathy; Phytotoxicity;
 Shoots; Allelopathins; Root systems; Plant morphology; Growth
 rate; Root hairs; Scanning electron microscopy
 
 Abstract:  The aqueous extract of alfalfa (Medicago sativa L.)
 shoots inhibits root elongation, shoot elongation, and/or
 germination of alfalfa itself--a phenomenon termed
 autotoxicity. The study of the mode of action of allelopathic
 and/or autotoxic compounds at the plant organ and cellular
 levels is limited by the depth of field, resolution, and
 magnification of a light microscope compared to a scanning
 electron microscope. Scanning electron microscopy techniques
 were used to study the morphology and anatomy of the roots
 inhibited by the water-extract of alfalfa shoots.
 Investigations on the morphology of shoot-aqueous-extract-
 inhibited, 5-d old 'WL-3l6' alfalfa roots revealed a 46%
 reduction in density and 54% reduction in length of root hairs
 compared to the control. Anatomical differences between the
 inhibited and uninhibited alfalfa roots were also observed.
 Shoot aqueous extract did not cause clogging of xylem vessels.
 The scanning electron microscope is a valuable tool in the
 study of the mode of action of allelopathic or autotoxic
 compounds at the plant organ as well as cellular levels.
 
 
 217                                NAL Call. No.: QD415.A1B5
 Seasonal patterns in the allelochemicals of Pseudotsuga
 menziesii, Picea engelmanii and Abies concolor.
 Wagner, M.R.; Clancy, K.M.; Tinus, R.W.
 Oxford : Pergamon Press; 1990.
 Biochemical systematics and ecology v. 18 (4): p. 215-220;
 1990.  Includes references.
 
 Language:  English
 
 Descriptors: Arizona; Pseudotsuga menziesii; Picea
 engelmannii; Abies concolor; Choristoneura occidentalis;
 Allelochemicals; Pest resistance; Seasonal fluctuations;
 Terpenoids; Tannins; Phenols; Nitrogen; Ratios; Leaves
 
 
 218                                NAL Call. No.: QD415.A1J6
 Seasonal patterns of juglone in soil beneath Juglans nigra
 (black walnut) and influence of J. nigra on understory
 vegetation.
 De Scisciolo, B.; Leopold, D.J.; Walton, D.C.
 New York, N.Y. : Plenum Press; 1990 Apr.
 Journal of chemical ecology v. 16 (4): p. 1111-1130; 1990 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Juglans nigra; Allelopathy; Soil analysis;
 Quinones; Seasonal variation; Undergrowth
 
 Abstract:  The allelopathic nature of J. nigra L. was
 investigated in several planted mixed hardwood stands located
 near Syracuse, New York. Concentrations of chloroform-
 extracted juglone from soil collected beneath J. nigra was
 determined by thin-layer chromatography (TLC) and high-
 pressure liquid chromatography (HPLC). Soil juglone
 concentrations were corrected based on recovery of synthetic
 juglone added to soil. Soil juglone levels were high in the
 spring, decreased during the summer, and were high again in
 the fall. The quantification of juglone from soil by HPLC was
 found to be more accurate than by TLC. Regression analysis
 indicated that individual tree variation in soil juglone
 levels could not be explained by differences in soil moisture,
 pH, organic matter content, and texture. The results of
 juglone recovery experiments suggest that chloroform-
 extractable juglone does not persist in soil. Juglone
 degradation by microorganisms could only explain a portion of
 the juglone decline. Ordinations revealed that the herbaceous
 and woody vegetation beneath J. nigra, in comparison to
 vegetation beneath Acer saccharum and Quercus rubra, is
 distinct in only one of the four stands studied. This
 vegetational difference did not appear to be a consequence of
 any strong allelopathic influences of J. nigra (Scheffe's
 method of contrast, chi-square analysis). The allelopathic
 nature of juglone under these field conditions is
 questionable.
 
 
 219                                   NAL Call. No.: 450 AN7
 Seminal root growth in sorghum (Sorghum bicolor) under
 allelopathic influences from residues of taro (Colocasia
 esculenta).
 Pardales, J.R. Jr; Kono, Y.; Yamauchi, A.; Iijima, M.
 London : Academic Press; 1992 Jun.
 Annals of botany v. 69 (6): p. 493-496; 1992 Jun.  Includes
 references.
 
 Language:  English
 
 Descriptors: Sorghum bicolor; Roots; Growth; Inhibition;
 Colocasia esculenta; Plant residues; Allelopathy;
 Phytotoxicity
 
 Abstract:  The length of the seminal root (SR) axis and the
 number and length of lateral roots (LRs) of sorghum (Sorghum
 bicolor Moench) were markedly inhibited by taro [Colocasia
 esculenta (L.) Schott] residues incorporated into a sand
 growing medium. The sand profile was divided equally into
 zones with and without residues. Production and elongation of
 the first-order LRs of the SR axis facing the zone containing
 taro residues were severely suppressed. On the side facing the
 zone that was free of residues, production and elongation of
 LRs was not inhibited. SR and LR growth was drastically
 impaired and many plants were killed when taro residues were
 incorporated in large amounts into the uppermost 2 cm of the
 growing medium. The activity of the allelopathic substances in
 the root zone appeared to be location-specific.
 
 
 220                                 NAL Call. No.: 450 J8224
 Short-term effects of ferulic acid on ion uptake and water
 relations in cucumber seedlings.
 Booker, F.L.; Blum, U.; Fiscus, E.L.
 Oxford : Oxford University Press; 1992 May.
 Journal of experimental botany v. 43 (250): p. 649-655; 1992
 May.  Includes references.
 
 Language:  English
 
 Descriptors: Cucumis sativus; Ferulic acid; Inhibition;
 Potassium; Nitrate; Ion transport; Ion uptake; Leaf water
 potential; Turgor; Roots; Seedlings; Allelopathy
 
 Abstract:  Ferulic acid (FA) is commonly found in soils and is
 considered an allelochemical. Studies have suggested that FA
 and other phenolic acids decrease plant growth in part by
 decreasing the absorption of mineral nutrients and water.
 However, no studies have examined these parameters in a single
 experimental system to investigate how FA affected both ion
 uptake and plant-water relations in whole plants. Using intact
 cucumber (Cucumis sativus L. cv. Early Green Cluster)
 seedlings, we examined short-term effects of FA on ion uptake
 kinetics, transport promoters and inhibitors, and water
 relations as indicated by a pressure-volume analysis. We found
 that after 3 h of treatment, 200 micromolar FA inhibited net
 ion uptake, particularly NO3(-1), and promoted net K+ efflux
 from seedling roots. The addition of fusicoccin, a K+
 transport promoter, counteracted the inhibitory effect of FA
 on net K+ uptake. Concurrent treatment of seedlings with FA
 and tetraethylammonium, a channel-blocking salt, reduced
 average K+ efflux by 66%. Treatment of seedlings with FA also
 decreased leaf water potential and turgor pressure (P(T)).
 However, decreased leaf water potential and P(T) were not
 caused by changes in the osmotic properties of the symplast or
 stomatal conductance. A decrease in water absorption is a
 likely explanation for the loss of P(T) observed. The results
 of our experiments indicate that both ion uptake and plant-
 water relations can be adversely affected by FA.
 
 
 221                                   NAL Call. No.: QK1.C83
 Significance of phenolic compounds in plant-soil-microbial
 systems. Siqueira, J.O.; Nair, M.G.; Hammerschmidt, R.; Safir,
 G.R. Boca Raton, Fla. : CRC Press; 1991.
 Critical reviews in plant sciences v. 10 (1): p. 63-121; 1991. 
 Literature review.  Includes references.
 
 Language:  English
 
 Descriptors: Plant physiology; Phenolic compounds;
 Biosynthesis; Metabolites; Soil chemistry; Allelopathins;
 Plant growth regulators; Host parasite relationships;
 Symbiosis; Literature reviews
 
 
 222                                NAL Call. No.: QD415.A1J6
 Soil transformation of 2(3H)-benzoxazolone of rye into
 phytotoxic 2-amino-3H-phenoxazin-3-one.
 Gagliardo, R.W.; Chilton, W.S.
 New York, N.Y. : Plenum Publishing Corporation; 1992 Oct.
 Journal of chemical ecology v. 18 (10): p. 1683-1691; 1992
 Oct.  Includes references.
 
 Language:  English
 
 Descriptors: Secale cereale; Allelopathy; Plant composition;
 Ketones; Metabolites; Soil biology; Phytotoxicity; Echinochloa
 crus-galli
 
 Abstract:  Nonsterile soil transforms the rye metabolite
 2(3H)-benzoxazolone (BOA) into 2-amino-3H-phenoxazin-3-one,
 which is an order of magnitude more toxic to barnyard grass
 than benzoxazolone. Benzoxazolone was recovered unchanged from
 sterile soil. However, o-aminophenol is converted to
 aminophenoxazinone by both sterile and nonsterile soil in the
 presence of air. Aminophenoxazinone is probably produced by
 microbial hydrolysis of benzoxazolone into o-aminophenol,
 which is oxidized to aminophenoxazinone in both sterile and
 nonsterile soil. No 2,2'-oxo-1,1'-azobenzene was found in any
 incubations of soil with benzoxazolone, o-aminophenol, or o-
 azophenol.
 
 
 223                                NAL Call. No.: 99.8 F7623
 Some future directions for biologically based vegetation
 control in forestry research.
 Jobidon, R.
 Ottawa : Canadian Institute of Forestry; 1991 Oct.
 The Forestry chronicle v. 67 (5): p. 514-519; 1991 Oct.  Paper
 presented at the symposium on "Recent Advances in Forest Pest
 Management", Oct 21, 1990, Sault Ste. Marie, Ontario. 
 Includes references.
 
 Language:  English
 
 Descriptors: Vegetation management; Weed control; Biological
 control; Phytotoxins; Allelopathy; Microbial pesticides;
 Research
 
 Abstract:  During the past decade, considerable research
 efforts have been devoted towards non-chemical weed control.
 Some of these efforts have been directed towards the control
 of forest weed species. Non-chemical control of forest
 vegetation encompasses many approaches and techniques and only
 a few of them are discussed in this paper. Three major and
 promising research areas are identified: 1) allelopathy, (2)
 microbially produced phytotoxins, and 3) bio-control. Each of
 these weed management strategies is briefly presented and
 discussed using examples from the forestry literature.
 
 
 224                                NAL Call. No.: QD415.A1J6
 Sorgoleone from root exudate inhibits mitochondrial functions.
 Rasmussen, J.A.; Hejl, A.M.; Einhellig, F.A.; Thomas, J.A. New
 York, N.Y. : Plenum Press; 1992 Feb.
 Journal of chemical ecology v. 18 (2): p. 197-207; 1992 Feb. 
 Includes references.
 
 Language:  English
 
 Descriptors: Sorghum bicolor; Root exudates; Allelopathy;
 Allelochemicals; Mitochondria; Zea mays; Glycine max; Electron
 transfer; Inhibition; Hydrophobicity; Weed control
 
 Abstract:  The aim of this investigation was to determine if
 sorgoleone (SGL), a hydrophobic compound in Sorghum bicolor
 (L.) Moench root exudate, interferes with mitochondrial
 functions. Tests were conducted on mitochondria isolated from
 etiolated soybean [Glycine max (L.) Merr.] and corn (Zea mays
 L.) seedlings. The data show SGL is a potent inhibitor of
 state 3 and state 4 respiration rates in both soybean and
 corn. Using either NADH, succinate, or malate as substrate,
 the I50 was about 0.5 micromolar SGL for state 3 and 5.0
 micromolar for state 4 based on 0.3-0.5 mg mitochondrial
 protein. Absorption spectra indicate SGL blocks electron
 transport at the b-c1 complex. These data show that disruption
 of mitochondrial function may be a mechanism of SGL-mediated
 growth inhibition previously reported and demonstrate a
 probable role of SGL in Sorghum allelopathy.
 
 
 225                                 NAL Call. No.: 450 P5622
 Soybean flavonoid effects on and metabolism by Phytophthora
 sojae. Rivera-Vargas, L.I.; Schmitthenner, A.F.; Graham, T.L.
 Oxford ; New York : Pergamon Press, 1961-; 1993 Mar.
 Phytochemistry v. 32 (4): p. 851-857; 1993 Mar.  Includes
 references.
 
 Language:  English
 
 Descriptors: Glycine max; Phytophthora; Plant pathogenic
 fungi; Host parasite relationships; Plant composition;
 Flavonoids; Growth; Plant development; Metabolism;
 Metabolites; Naringin; Growth inhibitors; Quercetin;
 Phytoalexins; Phytotoxicity; Disease resistance; Allelopathins
 
 Abstract:  Various soybean flavonoids were examined for their
 effects on the growth and development of Phytophthora sojae
 and for their metabolism by P. sojae. Three classes of
 molecules were identified based on their effects on growth.
 Coumestrol, biochanin A, genistein, naringenin and
 isorhamnetin were inhibitory at concentrations of 60-120
 micromolar and were fungicidal at 240 micromolar. Quercetin
 and its 3-O-beta-D-glucoside, isoquercitrin, caused
 significantly prolonged lags in P. sojae growth at 60-240
 micromolar, but were not fungicidal at any of these
 concentrations. Daidzein, formononetin, kaempferol, apigenin,
 chrysin and rutin were not inhibitory over this range. P.
 sojae rapidly hydrolysed all flavonoid glycosides tested and
 metabolized several flavonoids to non-aromatic products.
 However, the aglycones of the most inhibitory compounds were
 not significantly degraded. Metabolism of the compounds by P.
 sojae was a very early event and appeared to be associated
 predominantly with the hyphal tips. Several compounds had
 potentially interesting effects on fungal morphology and
 development. For example, genistein and its conjugates caused
 a marked swelling of the hyphal tip and an increase in the
 number of oogonia formed.
 
 
 226                                  NAL Call. No.: 450 J829
 Spatial patterning in plants: opposing effects of herbivory
 and competition. Bergelson, J.
 Oxford : Blackwell Scientific; 1990 Dec.
 Journal of ecology v. 78 (4): p. 937-948; 1990 Dec.  Includes
 references.
 
 Language:  English
 
 Descriptors: Washington; Poa annua; Senecio vulgaris;
 Deroceras reticulatum; Limax maximus; Plant competition;
 Allelopathy; Survival; Plant ecology; Spatial distribution;
 Herbivores; Feeding behavior
 
 Abstract:  (1) A field experiment was conducted to determine
 how the spatial distribution of Poa annua influenced the
 ability of Senecio vulgaris to establish in experimental
 plots. The relative effects of slug herbivores, dead
 individuals of Poa annua and live individuals of Poa annua on
 Senecio establishment were also investigated. (2) Senecio
 exhibited a higher rate of population growth when planted
 amidst clumped Poa than when planted amidst a random
 distribution of Poa. This change in population growth is due
 to increased survival of Senecio seedlings which emerge in
 areas having a low density of dead Poa. (3) The advantage
 associated with a clumped distribution of Poa was opposed by
 two other consequences of spatial distribution: herbivores
 consumed more Senecio seedlings. and intraspecific competition
 among Senecio seedlings was greater, where the grass was
 clumped than where it was randomly distributed. (4) Despite a
 number of direct effects, higher-order interactions and
 indirect effects, the net result of spatial patchiness is
 easily understood in this system because seedling suppression
 by dead Poa has overriding importance.
 
 
 227                                 NAL Call. No.: QL461.I57
 Status of biological control of Parthenium hysterophorus L. in
 India: a review.
 Srikanth, J.; Pushpalatha, N.A.
 Nairobi, Kenya : ICIPE Science Press; 1991 Aug.
 Insect science and its application v. 12 (4): p. 347-359; 1991
 Aug. Literature review.  Includes references.
 
 Language:  English
 
 Descriptors: India; Parthenium hysterophorus; Biological
 control; Weed control; Insects; Mites; Pathogens; Mycotoxins;
 Parasitic plants; Natural enemies; Allelopathy; Surveys;
 Literature reviews
 
 Abstract:  Biological control efforts on Parthenium
 hysterophorus L. (Asteraceae) in India have gained momentum
 after the limitations of other methods were realized. Native
 surveys revealed a large number of insects, but none of them
 was host specific. Although the introduced beetle Zygogramma
 bicolorata Pallister (Coleoptera: Chrysomelidae) has
 established at the sites of initial releases, its real impact
 on the weed and performance in different parts of the country
 need further evaluation. Fungal pathogens of the weed hold
 promise for classical as well as microherbicidal control. The
 use of parthenium phyllody MLO as a biocontrol agent requires
 establishment of host and vector specificity. Mycotoxins are a
 potential group of herbicides on which serious studies are yet
 to begin. Studies on control of the weed through interference
 and allelopathy by Cassia uniflora Mill.(= C. sericea Sw.)
 (Leguminosae) have produced promising results. Toxic leachates
 of C. uniflora and autotoxic principles of the weed deserve
 attention. integrated biocontrol strategies envisaged for
 wastelands using introduced insects and pathogens,
 allelopathic plants, and agroecosystems using native
 pathogens, mycotoxins and autotoxic principles, will help
 combat this apparently invincible weed.
 
 
 228                                 NAL Call. No.: 450 P5622
 Structure-activity relationships of phenylpropanoids as growth
 inhibitors of the green alga Selenastrum capricornutum.
 Della Greca, M.; Monaco, P.; Pollio, A.; Previtera, L.
 Oxford ; New York : Pergamon Press, 1961-; 1992 Dec.
 Phytochemistry v. 31 (12): p. 4119-4123; 1992 Dec.  Includes
 references.
 
 Language:  English
 
 Descriptors: Chlorophyta; Bioassays; Growth inhibitors;
 Propionic acid; Derivatives; Allelochemicals; Structure
 activity relationships; Molecular conformation
 
 Abstract:  Twenty-seven commercial or synthetic
 phenylpropanoids have been tested in broth against the
 unicellular alga Selenastrum capricornutum. The antialgal
 activity seems to be linked to the number as well as to the
 position of the methoxyl groups in the molecule. A slight
 effect of the side chain substitution was also observed.
 
 
 229                                NAL Call. No.: QD415.A1J6
 Sunflower aroma detection by the honeybee: study by coupling
 gas chromatography and electroantennography.
 Thiery, D.; Bluet, J.M.; Pham-Delegue, M.H.; Etievant, P.;
 Masson, C. New York, N.Y. : Plenum Press; 1990 Mar.
 Journal of chemical ecology v. 16 (3): p. 701-711; 1990 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Helianthus annuus; Plant composition; Aroma;
 Volatile compounds; Separation; Chemical analysis;
 Allelopathy; Interactions; Apis mellifera ligustica
 
 Abstract:  Combined electrophysiological recordings (EAG) and
 gas chromatographic separation were performed in order to
 investigate which volatile chemical components of a sunflower
 extract could be detected by honeybee workers and thus are
 likely to trigger the foraging behavior. A direct coupling
 device allowed for the stimulation of the antennal receptors
 with individual constituents of a polar fraction of the flower
 aroma shown to be attractive to bees. More than 100 compounds
 were separated from the extract. Twenty-four compounds
 elicited clear EAG responses. These compounds were identified
 by mass spectrometry (electronic impact and chemical
 ionisation). Both short- and long-chain aliphatic alcohols,
 one short-chain aliphatic aldehyde, one acid, two esters, and
 terpenic compounds were found to stimulate the antennal
 receptors. Six compounds identified in previous behavioral
 experiments were found to exhibit EAG activity. The chemicals
 screened by this method may be used for recognition of the
 plant odor and the selective behavior of honeybees.
 
 
 230                                 NAL Call. No.: QK900.J67
 Suppression of annuals by Tribulus terrestris in an abandoned
 field in the sandy desert of Kuwait.
 El-Ghareeb, R.M.
 Knivsta, Sweden : Opulus Press; 1991 Apr.
 Journal of vegetation science v. 2 (2): p. 147-154; 1991 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Kuwait; Tribulus terrestris; Weed competition;
 Allelopathy; Germination inhibitors; Invasion; Leachates;
 Phenolic compounds; Sandy soils; Annuals; Deserts; Plant
 communities; Site factors
 
 
 231                                   NAL Call. No.: 385 T29
 Synthesis and absolute configuration of lepidimoide, a high
 potent allelopathic substance from mucilage of germinated
 cress seeds. Kosemura, S.; Yamamura, S.; Kakuta, H.; Mizutani,
 J.; Hasegawa, K. Oxford : Pergamon Press; 1993 Apr16.
 Tetrahedron letters v. 34 (16): p. 2653-2656; 1993 Apr16. 
 Includes references.
 
 Language:  English
 
 Descriptors: Lepidium sativum; Seeds; Plant composition;
 Allelopathy; Disaccharides; Synthesis; Stereochemistry;
 Structure
 
 Abstract:  Lepidimoide (1) was isolated as a novel
 allelopathic substance, which promoted the shoot growth of
 different plant species but inhibited the root growth, from
 mucilage of germinated cress (Lepidium sativum L.) seeds. For
 example, lepidimoide (1) promoted the hypocotyl growth of
 etiolated Amaranthus caudatus L. at concentrations higher than
 3 micromolar and inhibited the root growth at concentrations
 higher than 100 micromolar. The growth-promoting activity in
 hypocotyls was 20 or 30 times as much as that of gibberellic
 acid. The structural study of lepidimoide, with spectral
 analyses and some chemical evidence, has showed that
 lepidimoide 1 is regarded as the uronic acid derivative
 bearing an alpha,beta-unsaturated carboxylate bonded to
 rhamnose via alpha-glucoside linkage. Thus, the intriguing
 structure as well as its unique biological activity prompted
 us to determine the absolute configuration of lepidimoide (1),
 by total synthesis. In this communication we wish to report a
 total synthesis and the absolute configuration of lepidimoide
 1.
 
 
 232                                   NAL Call. No.: 410 EC7
 Tannic acid, protein, and digestible carbohydrate: dietary
 imbalance and nutritional compensation in locusts.
 Raubenheimer, D.
 Tempe, Ariz. : The Society; 1992 Jun.
 Ecology : a publication of the Ecological Society of America
 v. 73 (3): p. 1012-1027; 1992 Jun.  Includes references.
 
 Language:  English
 
 Descriptors: Locusta migratoria; Diet studies;
 Allelochemicals; Carbohydrates; Digestibility; Interactions;
 Protein; Tannins
 
 
 233                                 NAL Call. No.: 64.8 C883
 Temperature stress and varietal resistance in rice: effects on
 whitebacked planthopper.
 Salim, M.; Saxena, R.C.
 Madison, Wis. : Crop Science Society of America; 1991 Nov.
 Crop science v. 31 (6): p. 1620-1625; 1991 Nov.  Includes
 references.
 
 Language:  English
 
 Descriptors: Oryza sativa; Cultivars; Genotypes; Varietal
 susceptibility; Sogatella furcifera; Pest resistance; Gene
 expression; Genotype environment interaction; Environmental
 temperature; Stress response; Heat stress; Cold stress; Plant
 composition; Chemical composition; Allelochemicals; Feeding
 behavior; Growth rate; Fecundity; Maturation period;
 Population dynamics; Host preferences
 
 Abstract:  Temperature greatly influences fundamental plant
 and pest physiological processes and interactions. This
 phytotron study sought to determine how optimum (29/21 degrees
 C), low (24/16 and 26/18 degrees C), and high (35/27 and 36/28
 degrees C) day/night temperature expression of regimes affect
 resistance of 'IR2035-117-3' (IR2035) and susceptibility of
 'Taichung Native 1' (TN1) rice (Oryza sativa L.) cultivars to
 the whitebacked planthopper, Sogatella furcifera (Horvath).
 Both low and high temperature stresses reduced plant growth
 and biomass, and influenced plant chemical composition. Insect
 food intake, growth, longevity, fecundity, and population
 increased significantly when IR2035 plants were grown at low
 and high (vs. optimum) temperature regimes (except 36/28
 degrees C). insect performance on temperature-stressed TN1
 plants was similar or superior to that on TN1 plants grown at
 optimum temperature, but at low temperatures, developmental
 period and longevity increased, while population decreased. At
 36/28 degrees C, the insect performed poorly on both resistant
 and susceptible plants. In spite of temperature-induced
 effects, the difference between resistance of IR2035 and
 susceptibility of TN1 remained distinct. Significantly more
 first instars died on TN1 plants painted with the steam-
 distillate extract of resistant IR2035 plants (grown at all
 temperatures studied) than on acetone-treated TN1 plants.
 Compared with unstressed plants, temperature-stressed plants
 had less allelochemicals. Since temperature-induced stresses
 enhance susceptibility of rice to insects, germplasm for S.
 furcifera resistance should be evaluated across a wide range
 of temperatures.
 
 
 234                                NAL Call. No.: QD415.A1J6
 Toxicity of allelopathic monoterpene suspensions on yeast:
 dependence on droplet size.
 Uribe, S.; Pena, A.
 New York, N.Y. : Plenum Press; 1990 Apr.
 Journal of chemical ecology v. 16 (4): p. 1399-1408; 1990 Apr. 
 Includes references.
 
 Language:  English
 
 Descriptors: Saccharomyces cerevisiae; Allelopathy;
 Monoterpenes; Emulsions; Toxicity; Droplet size
 
 Abstract:  The toxic effects of the allelopathic
 nonsubstituted monoterpenes beta-pinene and limonene on yeast,
 Saccharomyces cerevisiae, were proportional to the size of the
 monoterpene droplets in suspension. Both the toxic effects and
 the size of the droplets in suspension were decreased by
 adding different solvents with the monoterpene as follows:
 dimethylsulfoxide approximately dimethylformamide >> ethanol >
 dioxane. Oxygen consumption was inhibited about 80% by 1 mM
 beta-pinene added in dimethylsulfoxide but less than 10% when
 beta-pinene was added in dioxane. Parallel decreases in
 droplet size and toxic effects of either monoterpene were also
 induced by hydrating the monoterpene-dimethylformamide or
 monoterpene-dimethylsulfoxide before addition to yeast.
 Molecular aggregation may be a mechanism to potentiate the
 allelopathic properties of monoterpenes when these associate
 with diverse soil components.
 
 
 235                                 NAL Call. No.: 421 EN895
 Toxicity of the limonoid allelochemical cedrelone to noctuid
 larvae. Koul, O.; Isman, M.B.
 Dordrecht : Kluwer Academic Publishers; 1992 Sep.
 Entomologia experimentalis et applicata v. 64 (3): p. 281-287;
 1992 Sep. Includes references.
 
 Language:  English
 
 Descriptors: Mamestra configurata; Peridroma saucia; Larvae;
 Limonoids; Allelochemicals; Cedrela odorata; Growth
 inhibitors; Toona ciliata; Toxicity
 
 
 236                                NAL Call. No.: S605.5.O74
 The truth about companion planting.
 Tozer, E.
 Emmaus, Pa. : Rodale Press, Inc; 1992 Feb.
 Organic gardening v. 39 (2): p. 63-64, 66-67; 1992 Feb.
 
 Language:  English
 
 Descriptors: Companion crops; Allelopathy; Plant protection
 
 
 237                                 NAL Call. No.: 450 P5622
 Uptake and detoxification of salicylic acid by Vicia faba and
 Fagopyrum esculentum.
 Schulz, M.; Schnabl, H.; Manthe, B.; Schweihofen, B.; Casser,
 I. Oxford : Pergamon Press; 1993 May.
 Phytochemistry v. 33 (2): p. 291-294; 1993 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Vicia faba; Fagopyrum esculentum; Roots;
 Salicylic acid; Allelopathins; Uptake; Detoxification;
 Biosynthesis; Metabolites; Chemical composition; Enzyme
 activity
 
 Abstract:  Roots of Vicia faba and Fagopyrum esculentum showed
 a characteristic tri-phasic uptake of salicylic acid. During
 the first, short phase, absorption was unaffected by O2
 depletion, vanadate or cysteine, which suggests a diffusion of
 salicylic acid into the apoplast and some penetration into the
 cytoplasm. The second phase, continuing for about 6-8 (V.
 faba) and 5-8 hr (F. esculentum), was stationary and no
 obvious uptake was observed. The third phase was characterized
 by an active uptake. The absorbed salicylic acid was
 differently detoxified by the two species and the resulting
 compounds were identified. Vicia faba glucosylated salicylic
 acid to form o-beta-D-glucosylhydroxybenzoic acid, whereas F.
 esculentum oxidized it to 2,5-dihydroxybenzoic acid and
 glucosylated this product at the 5-OH group. The enzymes
 involved seemed to be induced by salicylic acid. Most of the
 detoxification occurred during the phase of active uptake.
 
 
 238                                   NAL Call. No.: QD1.A45
 Use of natural products in pest control: developing research
 trends. Hedin, P.A.
 Washington, D.C. : The Society; 1991.
 ACS Symposium series - American Chemical Society (449): p.
 1-11; 1991.  In the series analytic: Naturally occurring pest
 bioregulators / edited by P. A. Hedin.  Literature review. 
 Includes references.
 
 Language:  English
 
 Descriptors: Insect control; Biological control; Pheromones;
 Allelochemicals; Microbial pesticides; Weed control;
 Literature reviews
 
 
 239                                  NAL Call. No.: 442.8 Z8
 Variation within flax (Linum usitatissimum) and barley
 (Hordeum vulgare) in response to allelopathic chemicals.
 Ray, H.; Hastings, P.J.
 Berlin, W. Ger. : Springer International; 1992.
 Theoretical and applied genetics v. 84 (3/4): p. 460-465;
 1992.  Includes references.
 
 Language:  English
 
 Descriptors: Hordeum vulgare; Linum usitatissimum; Avena
 fatua; Linum; Genetic variation; Cultivars; Tolerance;
 Allelopathins; Allelopathy; P-coumaric acid; Shoots; Roots;
 Growth; Phenolic acids; Plant extracts
 
 Abstract:  A possible method of manipulating allelopathy would
 be to develop crop varieties showing an increased tolerance to
 allelopathic chemicals. We therefore examined four flax (Linum
 usitatissimum) varieties and two wild Linum species in the
 presence of p-coumaric acid and four barley (Hordeum vulgare)
 varieties in the presence of p-coumaric acid, scopoletin and
 wild oat (Avena fatua) extract. Analysis of variance indicates
 significant interaction between variety and treatment for
 shoot and root growth for seedling flax, shoot growth for
 older flax, and root growth for seedling barley. These
 differences in tolerance between varieties could be exploited
 to develop varieties with greater tolerances to the
 allelochemicals produced by weeds or in crop residues and
 therefore potentially more tolerant of the presence of weeds.
 
 
 240                                 NAL Call. No.: 1.98 AG84
 Victims no one mourns.
 Hays, S.M.
 Washington, D.C. : The Service; 1992 Feb.
 Agricultural research - U.S. Department of Agriculture,
 Agricultural Research Service v. 40 (2): p. 10-11; 1992 Feb.
 
 Language:  English
 
 Descriptors: Weed control; Aquatic weeds; Allelopathy;
 Biological control
 
 
 241                                 NAL Call. No.: 450 P5622
 Volatile compounds from leaves of Ceratiola ericoides by
 dynamic headspace sampling.
 Jordan, E.D.; Hsieh, T.C.Y.; Fischer, N.H.
 Oxford : Pergamon Press; 1992 Apr.
 Phytochemistry v. 31 (4): p. 1203-1208; 1992 Apr.  Includes
 references.
 
 Language:  English
 
 Descriptors: Florida; Empetraceae; Plant composition; Leaves;
 Volatile compounds; Chemical composition; Allelopathins;
 Seasonal variation
 
 Abstract:  Ceratiola ericoides is a shrub endemic to the
 Florida scrub community and has been investigated in
 conjunction with studies of allelopathic interactions that
 affect members of the adjacent sandhill community. Headspace
 volatiles of C. ericoides leaves collected in spring, summer
 and autumn were adsorbed on Tenax TA, thermally desorbed,
 cryogenically refocused, and identified by GC-MS. In spring
 leaves, hydrocarbons were most prevalent, while alcohols,
 aldehydes and ketones were most abundant in summer leaves.
 Esters were the major components in autumn leaves.
 
 
 242                                NAL Call. No.: QD415.A1J6
 Volatile seed germination inhibitors from plant residues.
 Bradow, J.M.; Connick, W.J. Jr
 New York, N.Y. : Plenum Press; 1990 Mar.
 Journal of chemical ecology v. 16 (3): p. 645-666; 1990 Mar. 
 Includes references.
 
 Language:  English
 
 Descriptors: Allium cepa; Daucus carota; Lycopersicon
 esculentum; Allelopathy; Volatile compounds; Germination
 inhibitors; Cover crops
 
 Abstract:  Volatile emissions from residues of the winter
 cover legumes, Berseem clover (Trifolium alexandrinum L.).
 hairy vetch [Vicia hirsuta (L.) S.F. Gray], and crimson clover
 (Trifolium incarnatum L.), inhibited germination and seedling
 development of onion, carrot. and tomato. Using GC-MS, 31 C2-
 C10 hydrocarbons, alcohols, aldehydes, ketones, esters,
 furans, and monoterpenes were identified in these residue
 emission mixtures. Mixtures of similar compounds were found in
 the volatiles released by herbicide-treated aerial and root
 residues of purple nutsedge (Cyperus rotundus L.) and the
 late-season woody stems and roots of cotton (Gossypium
 hirsutum L.). Vapor-phase onion, carrot. and tomato seed
 germination bioassays were used to determine the time- and
 concentration-dependent inhibition potential of 33 compounds
 that were either identified in the plant residue emissions or
 were structurally similar to identified compounds. Cumulative
 results of the bioassays showed that (E)-2-hexenal was the
 most inhibitory volatile tested, followed by nonanal, 3-
 methylbutanal, and ethyl 2-methylbutyrate. All the volatile
 mixtures examined contained at least one compound that greatly
 inhibited seed germination.
 
 
 243                                 NAL Call. No.: 450 P5622
 Volatiles from litter and soil associated with Ceratiola
 ericoides. Jordan, E.D.; Hsieh, T.C.Y.; Fischer, N.H.
 Oxford : Pergamon Press; 1993 May.
 Phytochemistry v. 33 (2): p. 299-302; 1993 May.  Includes
 references.
 
 Language:  English
 
 Descriptors: Florida; Empetraceae; Allelopathins; Plant
 composition; Volatile compounds; Litter (plant); Alcohols;
 Aldehydes; Ketones; Benzene; Derivatives; Terpenoids; Soil
 analysis
 
 Abstract:  Litter of Ceratiola ericoides and soil associated
 with this shrub were analysed for their volatiles by dynamic
 headspace sampling, followed by GC-MS identification. In
 litter volatiles, 1-octene, 3-octanol and 1-pentanol were most
 prevalent, while aliphatic alcohols and ketones were most
 abundant in soil. The major classes of volatiles, in both
 litter and soil, were aliphatic alcohols, aldehydes and
 ketones, and lesser amounts of benzenoids, monoterpenes and
 sesquiterpenes.
 
 
 244                                   NAL Call. No.: 100 AR42F
 Weed control with crop allelopathy.
 Dilday, R.H.; Frans, R.E.; Semidey, N.; Smith, R.J.; Oliver,
 L.R. Fayetteville, Ark. : The Station; 1992 Jul.
 Arkansas farm research - Arkansas Agricultural Experiment
 Station v. 41 (4): p. 14-15; 1992 Jul.  Includes references.
 
 Language:  English
 
 Descriptors: Oryza sativa; Helianthus annuus; Gossypium
 hirsutum; Glycine max; Weed control; Allelopathy; Crop yield
 
 
 
                          Author Index
 
 Abdel-Hady, N.F.  20
 Aerts, R.J.  37
 Ahmad, S.  71, 163
 Al-Dulaimy, S.M.  36
 Aletor, V.A.  15
 Aliotta, G.  196
 Allen, P.J.  186
 Alsaadawi, I.S.  36
 Altman, D.W.  79
 An, Z.Q.  209
 Anaya, A.L.  38, 109, 127, 188
 Anbu, D.A.  62
 Anderson, R.L.  91
 Anwar, T.  194
 Apt, W.J.  151
 Arnason, J.T.  84, 147
 Ash, J.E.  65
 Auld, D.L.  17
 Austin, D.F.  75
 Babu, R.C.  40
 Barbosa, P.  13, 14, 67, 154
 Barbour, M.G.  207
 Barnes, J.P.  46
 Bazzaz, F.A.  112
 Bean, G.A.  11
 Benedict, J.H.  144
 Berenbaum, M.R.  85
 Bergelson, J.  226
 Bernard, C.B.  147
 Bewick, T.A.  60
 Bhatt, B.P.  189
 Bluet, J.M.  229
 Blum, M.S.  157
 Blum, U.  18, 116, 117, 120, 126, 153, 158, 187, 220
 Boerner, R.E.J.  61
 Boethel, D.J.  128
 Booker, F.L.  220
 Bowers, M.D.  112
 Boyd, R.S.  207
 Bradbury, J.H.  86
 Bradow, J.M.  203, 242
 Brattsten, L.B.  72
 Bray, R.O.  155
 Brede, A.D.  136
 Bremner, J.M.  121
 Brown, P.D.  17
 Bruin, J.  192
 Bullock, D.G.  92
 Burden, B.J.  213
 Caboun, Vladimir  8
 Calera, M.R.  38
 Camps, F.  161
 Caporale, G.  195
 Cappelletti, E.M.  195
 Capua, S.  148
 Carballeira, A.  206
 Carino, F.A.  144
 Casser, I.  237
 Cast, K.G.  16
 Castaneda, P.  127
 Caswell, E.P.  151
 Chase, W.R.  3, 4
 Chauhan, D.S.  189
 Chen, P.K.  191
 Chilton, W.S.  222
 Choesin, D.N.  61
 Chou, C.H.  23, 43, 210
 Clancy, K.M.  217
 Codella, S.G. Jr  110
 Cohen, E.  148
 Coll, J.  161
 Colwell, A.  107
 Connick, W.J. Jr  242
 Corcuera, L.J.  70
 Cowgill, U.M.  81
 Craig, R.  134
 Crooks, J.R.  123
 Cruz Ortega, R.  109
 Dahl, B.E.  132
 Dakshini, K.M.M.  139, 143, 165, 166
 De Scisciolo, B.  218
 Deb, P.R.  145
 DeFrank, J.  151
 Della Greca, M.  1, 228
 Dick, W.A.  52
 Dicke, M.  192
 Dilday, R.H.  244
 Doolittle, R.E.  169
 Dornbos, D.L. Jr  175
 Dowd, P.F.  90, 97, 162
 Downum, K.R.  150
 Duke, S.O.  181
 Dusky, J.A.  60
 Dussourd, D.E.  138
 Dyck, E.  93
 Effects of some compounds isolated from Celaenodendron
 mexicanum Standl (Euphorbiaceae) on seeds and phytopathogenic
 fun  127
 Einhellig, F.A.  118, 190, 224
 El Abdaoui, F.  122
 El-Darier, S.M.  179
 El-Ghareeb, R.M.  230
 Elakovich, S.D.  19, 39
 Elissalde, M.H.  79
 Eller, F.J.  174
 Ells, J.E.  29
 Elmore, C.D.  75
 Elmore, C.L.  199
 Etievant, P.  229
 Faeth, S.H.  105
 Fajer, E.D.  112
 Faleiro, L.J.  150
 Farkas, P.  137
 Fengyou, W.  27
 Ferguson, D.E.  41
 Ferriss, R.S.  209
 Fischer, N.H.  241, 243
 Fiscus, E.L.  220
 Fogal, W.H.  110
 Fogleman, J.C.  167
 Foy, C.L.  122
 Frank, M.R.  167
 Frans, R.E.  244
 Fujikura, J.  173
 Fukuhara, K.  171
 Fulbright, N.  34, 141
 Fulbright, T.E.  34, 141
 Gabel, B.  137
 Gadgil, R.L.  186
 Gagliardo, R.W.  222
 Galindo, J.C.G.  197
 Gallagher, S.S.  186
 Gallardo, F.  128
 Garcia, M.R.  127
 Gavilanes-Ruiz, M.  109
 Gerig, T.M.  18, 120, 187
 Gerson, U.  148
 Ghazi, M.  6
 Gilbert, H.  56, 57
 Goel, U.  87, 205
 Gopal, B.  87
 Gould, F.  100, 208
 Graham, T.L.  225
 Graves, C.H. Jr  146
 Grodzinskii, A. M.  58
 Gross, P.  13, 14, 154
 Grossman, J.  77
 Guo, B.Z.  209
 Hammer, B.C.  86
 Hammerschmidt, R.  221
 Harrison, H.F. Jr  102, 131, 172
 Hartung, A.C.  168
 Hasegawa, K.  53, 170, 231
 Hastings, P.J.  239
 Hays, S.M.  240
 Heath, R.R.  169, 184
 Hedin, P.A.  76, 115, 119, 129, 146, 204, 238
 Hegazy, A.K.  20
 Hegde, R.S.  47, 89, 216
 Heisey, R.M.  21, 130
 Hejl, A.M.  118, 224
 Hendrix, J.W.  209
 Hernandez, B.E.  127
 Hernandez-Bautista, B.E.  188
 Hilton, A.S.  42
 Hodgson, E.  100, 208
 Hogan, M.E.  54, 101, 114
 Hogberg, P.  27
 Holappa, L.D.  18, 116
 Hradsky, P.  137
 Hsieh, T.C.Y.  241, 243
 Hurst, H.R.  75
 Iijima, M.  219
 Inderjit  139, 143, 165, 166
 Innocenti, G.  195
 Inoue, M.  12
 Isenhour, D.J.  9
 Isman, M.B.  176, 235
 Iyengar, S.  84
 Jabbar, A.  194
 Jakubas, W.J.  211
 Jarvis, B.B.  11
 Jenkins, J.N.  115, 129, 204
 Jermy, T.  200
 Jimenez Estrada, M.  109
 Jimenez-Estrada, M.  188
 Jobidon, R.  223
 Jones, G.P.D.  104
 Jordan, E.D.  241, 243
 Joshi, S.  73
 Kakuta, H.  231
 Kalburtji, K.L.  125
 Kaspar, T.C.  111
 Kelsey, R.G.  155
 Kemper, J.  154
 Keogh, D.P.  123
 Kester, K.M.  67
 Khalique, F.  194
 Khara, A.  30
 Kholdebarin, B.  32, 33
 Kil, B.S.  35, 64, 149
 King, L.D.  18, 187
 Klein, K.  126, 158, 187
 Koide, R.T.  180
 Komai, K.  83
 Konno, T.  208
 Kono, Y.  219
 Kosemura, S.  170, 231
 Koul, O.  235
 Krenzer, E.G. Jr  16
 Kubo, I.  171
 Kuti, J.O.  11
 Laird, D.W.  146
 Lam, J.  147
 Landenberger, B.D.  81
 Lawrence, J.G.  107
 Lawrence, P.O.  69
 Leather, G.R.  191
 Lee, S.Y.  149
 Lee, Y.F.  23
 Leopold, D.J.  218
 Leu, L.L.  43
 Levi, P.  208
 Levi, P.E.  100
 Lewis, W.J.  142, 174
 Li, H.H.  12, 53
 Li, M.  180
 Liebman, M.  93
 Lindroth, R.L.  140
 Liu, D.L.  51
 Lloret, F.  202
 Lovett, J.V.  51, 80, 82
 Lydon, J.  181
 Lyu, S.W.  117, 187
 Macias, F.A.  182, 197
 Mangoni, L.  1
 Manner, G.D.  101
 Manners, G.D.  54, 114
 Mansour, K.S.  20
 Manthe, B.  124, 237
 Manthe, Barbara,  44
 Marion-Poll, F.  137
 Martin, V.L.  52
 Mason, J.R.  211
 Massanet, G.M.  197
 Masson, C.  229
 Mata, R.  38, 127
 Matizha, W.  132
 May, F.E.  65
 McCaffrey, J.P.  17
 McCarty, G.W.  121
 McCarty, J.C. Jr  119
 McCoy, E.L.  52
 McCrady, J.J.  42
 McPherson, J.K.  16
 McSay, A.E.  29
 Meijden, E. van der  37
 Meissner, R.  26
 Miles, P.W.  185
 Miller, D.A.  47, 89, 216
 Miller, H.G.  63
 Miller, R.W.  175
 Milman, I.A.  7
 Mishra, S.K.  3
 Mitchell, E.R.  184
 Mitchell, M.J.  123
 Mizutani, J.  12, 53, 55, 170, 173, 231
 Moellenbeck, D.J.  5
 Mohamed, M.A.  5
 Mokhtari-Rejali, N.  11
 Molina, A.  206
 Molinaro, A.  1
 Molinillo, J.M.G.  182
 Monaco, P.  1, 196, 228
 Morand, P.  84
 Moroz, P. A.  59
 Morra, M.J.  17
 Mosjidis, J.A.  125
 Mossler, M.A.  60
 Muehleisen, D.P.  144
 Mumma, R.O.  134
 Munesada, K.  74
 Murphy, S.D.  96
 Murray, D.S.  88, 98
 Myers, G.A.  6
 Myster, R.W.  106
 Nair, M.G.  2, 3, 4, 46, 168, 221
 Nathawat, G.S.  205
 Nel, P.C.  26
 Netzly, D.  108
 Nilsson, M.C.  27, 28, 160
 Nishimoto, R.K.  83
 Nishimura, H.  12, 53, 55
 Norman, J.O.  79
 Norris, D.M.  213
 Nowbahari, B.  94
 Obee, E.M.  159
 Oertli, J.J.  32
 Ogawa, H.  173
 Oguntimein, B.O.  19
 Oliver, L.R.  244
 Ormeno-Nunez, J.  99, 214
 Pacheco, D.Y.  13
 Pardales, J.R. Jr  219
 Pardini, R.S.  163, 212
 Parrott, W.L.  115, 129, 204
 Pastore, J.  212
 Peirce, L.C.  63
 Pena, A.  234
 Peng, Z.  185
 Pereda-Miranda, R.  38
 Perez, F.J.  25, 99, 214
 Perumal, R.K.P.  40
 Peterson, J.K.  102, 131, 172
 Pham-Delegue, M.H.  229
 Philogene, B.J.R.  84, 147
 Pickett, S.T.A.  106
 Pinto, G.  196
 Plapp, F.W. Jr  144
 Pollard, A.J.  16
 Pollio, A.  196, 228
 Posthumus, M.A.  192
 Previtera, L.  1, 196, 228
 Prevost, G.  142
 Pritsos, C.A.  212
 Provan, G.J.  13, 14
 Proveaux, A.T.  169
 Puri, S.  30
 Purvis, C.E.  103, 104
 Pushpalatha, N.A.  227
 Putnam, A.R.  2, 3, 4, 46, 168
 Quisenberry, S.S.  5
 Raffa, K.F.  110
 Rahman, A.  152
 Rani, M.S.  40
 Rao, A.S.  215
 Rasmussen, J.A.  118, 224
 Raubenheimer, D.  232
 Ray, H.  239
 Reigosa, M.J.  206
 Reinhardt, C.F.  26
 Reynolds, S.E.  66
 Riffle, M.S.  88, 98
 Rivera-Vargas, L.I.  225
 Rizvi, S. J. H., 1955  50
 Rizvi, V.,  50
 Rose, R.L.  100, 208
 Rosenthal, G.A.  177
 Sabelis, M.W.  192
 Safir, G.R.  221
 Sakeri, F.A.K.  36
 Salim, M.  183, 233
 Sanchez Nieto, S.  109
 Sandberg, A.M.  186
 Saric, Taib  48
 Saxena, R.C.  183, 233
 Schenk, S.U.  68
 Schloman, W.W. Jr  42
 Schmidt, E.L.  121
 Schmitthenner, A.F.  225
 Schnabl, H.  124, 237
 Schneider, D.  133
 Schreiber, M.M.  156
 Schulz, M.  124, 237
 Schutt, C.  108
 Schweihofen, B.  237
 Semidey, N.  244
 Sexton, O.J.  107
 Sgaramello, R.P.  98
 Shafer, S.R.  153
 Shakeel, M.A.  194
 Shen, S.K.  90, 97
 Sheriff, M.M.  40
 Shilling, D.G.  60
 Siddiqui, H.L.  74
 Singh, D.B.  113
 Siqueira, J.O.  221
 Slansky, F. Jr  78, 135
 Smirle, M.J.  176
 Smith, A.E.  49, 198
 Smith, R.J.  244
 Smith, S.L.  123
 Snoeijer, W.  37
 Snook, M.E.  9
 Souza, I.F.  190
 Spencer, G.F.  175
 Srikanth, J.  227
 Stermitz, F.R.  13, 14
 Stevens, K.L.  45
 Stipanovic, R.D.  79
 Suchy, V.  137
 Suga, T.  74
 Sullia, S.B.  62
 Suresh, K.K.  31
 Swain, L.A.  150
 Swaminathan, C.  31
 Tahir, S.  194
 Takabayashi, J.  10, 192
 Takahashi, S.  10
 Tang, C.S.  83, 151
 Teasdale, J.R.  164
 Tesar, M.B.  95
 Thibout, E.  94
 Thiery, D.  137, 229
 Thomas, J.A.  224
 Timmins, W.A.  66
 Tingle, F.C.  184
 Tinus, R.W.  217
 Todaria, N.P.  189
 Torres, A.  182
 Torres, B.A.  127
 Tozer, E.  236
 TSentral  58
 Tumlinson, J.H.  169, 174
 Turlings, T.C.J.  169, 174
 Uribe, S.  234
 Varela, R.M.  182
 Velasco-Ibarra, L.  188
 Verpoorte, R.  37
 Vinaya Rai, R.S.  31
 Waddell, T.  147
 Wagner, M.R.  217
 Wahab, Z.B.  111
 Waller, G.R.  16, 88, 98
 Walters, D.S.  134
 Walton, D.C.  218
 Wambolt, C.L.  155
 Wardle, D.A.  152
 Weidenhamer, J.D.  159
 Weinhold, L.C.  163
 Weisbrod, A.V.  140
 Wentworth, T.R.  187
 Werner, D.  68
 Werstiuk, N.H.  84
 Weyman-Kaczmarkowa, W.  178
 Wheeler, G.S.  78, 135
 Whitenack, C.J.  2
 Williams, L. III  17
 Williamson, G.B.  159
 Wilson, R.L.  9
 Wink, M.  133, 201
 Wiseman, B.R.  9
 Witte, L.  201
 Wojcik-Wojtkowiak, D.  178
 Wooten, J.W.  39
 Worsham, A.D.  18, 22, 187
 Xavier, A.  24
 Yamamura, S.  170, 231
 Yamane, A.  55, 173
 Yamauchi, A.  219
 Yu, S.J.  193
 Yun, K.W.  35, 64, 149
 Zackrisson, O.  27, 28, 160
 Zedler, P.H.  202
 
 
                          Subject Index
 
 2-tridecanone  100
 Abies concolor  217
 Abscisic acid  116
 Abutilon theophrasti  102, 164, 175, 180
 Acacia leucophloea  31
 Acer rubrum  13
 Acinetobacter calcoaceticus  3
 Acremonium coenophialum  209
 Acrolepiopsis assectella  94
 Activated carbon  60
 Adaptability  138
 Adaptation  67
 Adenosinetriphosphatase  109
 Adverse effects  60
 Aedes aegypti  123
 Aegilops cylindrica  91
 Aeration  52
 Agroforestry  189
 Agrostis stolonifera var. palustris  136
 Ailanthus altissima  21, 107, 130
 Ajuga  161
 Alanine  68
 Alcohols  243
 Aldehydes  204, 243
 Aldrin  148
 Algae  196
 Alkaloids  6, 13, 37, 133, 154, 167, 209
 Allelochemicals  3, 4, 9, 10, 11, 12, 16, 17, 51, 53, 66, 67,
 69, 70, 71, 72, 76, 78, 80, 81, 82, 85, 88, 89, 90, 94, 97,
 101, 105, 110, 111, 115, 119, 122, 123, 129, 133, 137, 138,
 140, 142, 147, 148, 150, 154, 155, 157, 161, 162, 165, 167,
 174, 176, 178, 182, 183, 184, 185, 192, 193, 194, 196, 200,
 201, 204, 208, 213, 217, 224, 228, 232, 233, 235, 238
 Allelopathic agents  48, 50, 58, 59
 Allelopathins  19, 21, 24, 28, 37, 41, 47, 56, 57, 61, 62, 65,
 68, 74, 84, 96, 102, 113, 128, 134, 145, 146, 170, 171, 177,
 197, 205, 209, 214, 216, 221, 225, 237, 239, 241, 243
 Allelopathy  1, 2, 3, 5, 6, 7, 8, 11, 12, 18, 19, 20, 21, 22,
 23, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 38, 39,
 40, 42, 43, 46, 47, 48, 49, 50, 51, 52, 53, 55, 56, 57, 58,
 60, 63, 64, 65, 73, 75, 77, 80, 81, 82, 83, 87, 91, 92, 93,
 95, 98, 99, 102, 103, 104, 106, 107, 108, 109, 111, 114, 116,
 117, 118, 120, 121, 124, 125, 126, 127, 130, 131, 132, 136,
 139, 141, 143, 149, 151, 152, 153, 156, 158, 159, 160, 164,
 165, 166, 168, 170, 172, 173, 175, 178, 179, 180, 181, 186,
 187, 188, 189, 190, 191, 195, 198, 199, 202, 203, 206, 207,
 209, 210, 215, 216, 218, 219, 220, 222, 223, 224, 226, 227,
 229, 230, 231, 234, 236, 239, 240, 242, 244
 Allium  94
 Allium cepa  203, 242
 Allium sativum  148
 Allomones  15
 Allyl isothiocyanate  61
 Alocasia macrorrhiza  86
 Alpha-tomatine  128
 Alternaria brassicae  113
 Alternative farming  77, 80
 Amaranthus caudatus  170
 Amaranthus leucocarpus  38, 188
 Amaranthus retroflexus  102
 Ambrosia artemisiifolia  180
 Amino acids  24, 115
 Ammonia  121
 Ammonium  32, 33
 Ammonium nitrogen  32
 Anabaena flos-aquae  1
 Anacardic acid  134
 Analogs  144
 Animal behavior  142
 Annual habit  75
 Annuals  230
 Antagonism  4
 Antennaria microphylla  54, 101, 114
 Anthemis cotula  198
 Anthraquinones  12
 Antibiotics  62
 Antibodies  9
 Anticarsia gemmatalis  78, 135, 193
 Antifeedants  13, 14, 66, 154, 157, 161, 185, 200, 211
 Antifungal agents  62
 Antifungal properties  68, 108
 Antinutritional factors  15
 Antioxidants  71
 Apanteles  10
 Apis mellifera ligustica  229
 Apium graveolens  60
 Apparatus  136
 Application rates  22
 Application to land  42
 Aquatic plants  87
 Aquatic weeds  1, 240
 Arachis hypogaea  62, 122
 Arginine  177
 Arizona  217
 Aroma  229
 Artemisia annua  191
 Artemisia princeps  35, 64, 149
 Artemisia tridentata  155
 Asexual reproduction  75
 Asparagus officinalis  63, 168
 Aspergillus  145
 Astragalus  81
 Aucubin  112
 Australia  62
 Australian capital territory  65
 Avena fatua  25, 214, 239
 Avena sativa  25
 Azadirachtin  66
 Azo compounds  2
 Baccharis cordifolia  11
 Baccharis megapotamica  11
 Bark  21, 30, 65, 189
 Barriers  70
 Benzene  243
 Bibliographies  56, 57
 Binding site  144
 Bioassays  4, 5, 12, 16, 18, 19, 20, 35, 36, 38, 39, 47, 55,
 64, 65, 79, 96, 99, 102, 109, 127, 139, 143, 146, 159, 168,
 173, 175, 191, 197, 198, 228
 Biological activity in soil  32
 Biological control  10, 14, 17, 25, 36, 51, 67, 73, 75, 77,
 80, 98, 99, 115, 129, 131, 152, 169, 184, 186, 187, 192, 199,
 200, 211, 223, 227, 238, 240
 Biological control agents  73, 80, 154
 Biosynthesis  61, 134, 177, 215, 221, 237
 Boreal forests  28
 Brassica  77
 Brassica napus  17, 61
 Brassica nigra  113
 Bromus catharticus  152
 Cabt  18, 43, 91, 95, 127, 207
 Caffeic acid  121
 Caffeine  78
 Calcium  179
 California  202, 207
 Callus  54, 101, 114, 149
 Camellia  78
 Camellia sinensis  32, 33
 Canavanine  177
 Carbohydrates  232
 Carbon dioxide  112
 Carduus nutans  152
 Carrier proteins  144
 Carthamus tinctorius  91
 Carya pecan  146
 Cassia  73, 141
 Cassia occidentalis  102
 Casuarina equisetifolia  31
 Catechin  185
 Catechol oxidase  185
 Catharanthus roseus  37
 Cedrela odorata  235
 Cell cultures  33
 Cell division  20
 Cell suspensions  54, 101, 114
 Cenchrus ciliaris  34, 141
 Cenchrus incertus  132
 Chaparral  202
 Characterization  89
 Chemical analysis  68, 76, 96, 229
 Chemical composition  16, 74, 81, 83, 98, 99, 105, 134, 168,
 171, 188, 195, 233, 237, 241
 Chemical constituents of plants  1, 56, 215
 Chemical control  75, 91, 156
 Chemical ecology  43, 71, 105
 Chemical reactions  32
 Chemoreceptors  200, 211
 Chenopodium album  164
 Chitin  151
 Chlorella vulgaris  1
 Chloris gayana  151
 Chlorophyta  228
 Chloroplasts  118
 Choristoneura occidentalis  217
 Chromatography  171
 Chromolaena  19
 Cinchona  37
 Cinnamic acid  63, 159
 Cirsium vulgare  152
 Cladosporium  146
 Coastal plant communities  202, 207
 Coffea  78
 Cold stress  233
 Coleoptiles  214
 Colinus Virginianus  34
 Colocasia esculenta  219
 Colorado  91
 Community ecology  106
 Companion crops  236
 Competition  144
 Competitive ability  73, 75, 102, 131, 136, 175
 Compositae  84, 147
 Continuous cropping  95
 Control methods  80
 Cornus florida  13
 Cortaderia selloana  186
 Corticium rolfsii  145
 Cost benefit analysis  80
 Costa Rica  150
 Cotesia  67, 154
 Cotesia marginiventris  169, 174
 Cotyledons  32, 33
 Coumaric acids  153
 Coumarins  85, 195
 Cover crops  22, 77, 92, 164, 242
 Crop damage  63, 174
 Crop establishment  132
 Crop growth stage  111
 Crop losses  111
 Crop plants as weeds  25
 Crop production  80
 Crop residues  46, 47, 52, 89, 91, 111
 Crop weed competition  93, 102, 131, 136, 152, 165, 198
 Crop yield  95, 111, 131, 156, 244
 Cropping systems  156, 210
 Crops  46, 77
 Crotalaria juncea  151
 Crotalaria medicaginea  205
 Cruciferae  20, 207
 Cucumis sativus  4, 29, 116, 117, 120, 126, 153, 158, 220
 Cultivars  5, 26, 125, 183, 204, 233, 239
 Cultural control  77
 Cultural weed control  91, 93
 Culture filtrates  113, 145
 Cuscuta  201
 Cycling  179
 Cynodon dactylon  5, 36, 125
 Cyperus esculentus  131
 Cyperus rotundus  83
 Cyrtosperma chamissonis  86
 Cytochrome p-450  85, 100, 123, 167, 208
 Dactylis glomerata  152
 Daucus carota  148, 203, 242
 Decomposition  52
 Defense  94, 105, 192, 195
 Defense mechanisms  70, 138, 201
 Defoliation  105
 Degradation  17
 Delonix regia  43
 Density  95
 Derivatives  68, 159, 228, 243
 Deroceras reticulatum  226
 Desert plants  20
 Deserts  179, 230
 Design  136
 Desmanthus  141
 Desmodium  184
 Detoxification  85, 90, 97, 140, 157, 185, 237
 Developmental stages  179
 Diapause  69
 Dichanthium annulatum  34, 141
 Diet studies  232
 Diets  128, 135
 Digestibility  232
 Digestive tract  185
 Digitaria decumbens  151
 Dioscorea alata  86
 Dioscorea esculenta  86
 Diprion similis  110
 Disaccharides  231
 Disease resistance  108, 146, 177, 225
 Displacement  144
 Diterpenes  74, 161
 Diuraphis  70
 Drechslera  113
 Droplet size  234
 Drosophila  167
 Drosophila melanogaster  123
 Dry matter  111
 Dry matter accumulation  27
 Duneland plants  207
 Eating rates  78
 Echinochloa crus-galli  38, 188, 222
 Eclipta alba  102
 Ecosystems  167, 207, 210
 Ectomycorrhizas  27
 Egypt  179
 Electron transfer  224
 Electrophysiology  200
 Eleusine indica  102
 Emergence  63
 Empetraceae  241, 243
 Empetrum  27, 28, 160
 Emulsions  234
 Endophytes  209
 Environmental factors  70, 107, 132, 210
 Environmental temperature  233
 Enzyme activity  84, 100, 109, 123, 140, 148, 163, 237
 Epilachna varivestis  213
 Epoxides  148
 Eragrostis curvula  132
 Erosion  92
 Essential oils  19, 83, 88, 98, 149
 Establishment  164
 Esterases  140
 Eucalyptus globulus  65, 206
 Eucalyptus macrorhyncha  65
 Eucalyptus maculata  65
 Eucalyptus rossii  65
 Eucalyptus rubida  65
 Eucalyptus tereticornis  30, 31
 Eugenia uniflora  19
 Eupatorium capillifolium  198
 Euphorbia esula  54, 101, 114
 Euphorbia prostrata  36
 Euphorbiaceae  127
 Excretion  176
 Extraction  18
 Extracts  35, 54, 137, 143, 168
 Exudates  173
 Fagopyrum esculentum  237
 Fallow  95
 Fallow systems  187
 Farming systems  80
 Fat body  144
 Fatty acids  16, 134
 Feces  176
 Fecundity  110, 183, 233
 Feeding  135, 140
 Feeding behavior  5, 67, 78, 110, 138, 155, 213, 226, 233
 Feeding preferences  133
 Feeds  15
 Ferulic acid  116, 117, 121, 122, 126, 153, 158, 220
 Festuca arundinacea  152, 209
 Field experimentation  136
 Field tests  200
 Fire ecology  202
 Fixation  33
 Flavones  127
 Flavonoids  81, 90, 123, 129, 139, 213, 215, 225
 Florida  241, 243
 Flowers  43, 137
 Fodder crops  49
 Foliage  110
 Food consumption  135
 Food crops  189
 Forest ecology  8
 Forest litter  28, 65
 Forest trees  189
 Formica  94
 Formica fusca  94
 Fruit  171
 Fruit trees  59
 Fruits  195
 Fungi  162
 Fungicidal properties  146
 Fusarium  63, 145
 Gallic acid  122
 Gene expression  233
 Genetic engineering  72
 Genetic variation  61, 140, 239
 Genista  201
 Genotype environment interaction  233
 Genotypes  61, 112, 233
 Geographical distribution  83
 Geographical races  62
 Georgia  49
 Germination  18, 98
 Germination inhibitors  24, 37, 41, 42, 73, 102, 141, 165,
 175, 230, 242
 Germinationinhibitors  68
 Glabromicroplitis croceipes  142
 Gleichenia japonica  74
 Glomus etunicatum  180
 Glomus macrocarpum  209
 Glomus mosseae  209
 Glucose  33
 Glucosinolates  17
 Glutathione  163
 Glutathione transferase  148, 193
 Glycine max  18, 118, 145, 187, 189, 213, 224, 225, 244
 Glycoalkaloids  171
 Glycosides  38, 74, 140
 Glyphosate  95
 Golf green soils  136
 Gossypium  79, 184
 Gossypium arboreum  129
 Gossypium hirsutum  98, 115, 119, 204, 244
 Gossypol  115, 144
 Gramine  51
 Gramineae  68
 Grasslands  23
 Graviperception  215
 Green manures  77
 Greenhouse culture  60, 159
 Growing media  29
 Growth  9, 12, 16, 27, 31, 52, 55, 60, 104, 110, 111, 124,
 126, 135, 147, 159, 160, 170, 204, 205, 219, 225, 239
 Growth analysis  158
 Growth inhibitors  37, 41, 53, 61, 63, 64, 66, 68, 74, 107,
 113, 145, 165, 171, 175, 196, 214, 225, 228, 235
 Growth rate  6, 21, 24, 25, 47, 128, 131, 146, 152, 215, 216,
 233
 Growth retardants  205
 Growth retardation  146
 Growthpromoters  205
 Habitats  207
 Hawaii  151
 Heat stress  233
 Height  111
 Helianthus annuus  182, 229, 244
 Helicoverpa zea  9, 184, 193
 Heliothis subflexa  184
 Heliothis virescens  79, 84, 100, 115, 129, 184, 193, 204, 208
 Heliothis zea  144
 Herbicidal properties  21, 46, 181, 214
 Herbicides  2, 22, 156
 Herbivores  207, 226
 Heritability  142
 Histopathology  6
 Holcus lanatus  152
 Honeydew  185
 Hordenine  51, 154
 Hordeum  70
 Hordeum vulgare  51, 113, 239
 Host parasite relationships  67, 69, 154, 174, 183, 221, 225
 Host plants  167, 192
 Host preferences  233
 Host range  110
 Host-seeking behavior  169
 Hosts of plant pests  78
 Humus  28
 Hybrids  111, 134
 Hydrolases  97
 Hydrophobicity  224
 Hydroquinone  101, 114
 Hydroxamic acids  18, 25
 Hypocotyls  170
 Idaho  41
 Identification  89
 Ilex opaca  14
 Illinois  47
 Immobilization  32, 33, 91
 Imperata cylindrica  166
 In vitro  146
 Incorporation  47, 60
 India  62, 165, 205, 227
 Indiana  156
 Indigofera  205
 Induction  70, 100
 Infectivity  209
 Ingestion  157, 185
 Inhibition  12, 20, 24, 52, 55, 104, 109, 111, 125, 131, 139,
 152, 159, 166, 172, 195, 209, 219, 220, 224
 Insect attractants  174
 Insect control  5, 10, 14, 17, 94, 98, 115, 129, 134, 137,
 161, 169, 184, 192, 193, 194, 200, 208, 213, 238
 Insect pests  69, 70, 72, 92, 134, 150, 157, 161, 194
 Insecticidal action  123, 147, 194
 Insecticidal plants  194
 Insecticidal properties  76
 Insecticide resistance  72
 Insecticides  162
 Insects  227
 Integrated control  91
 Integrated pest management  82, 200
 Interactions  10, 87, 115, 229, 232
 Intercropping  93
 Introduced species  141
 Inula  7
 Invasion  207, 230
 Ion transport  220
 Ion uptake  220
 Ipomoea  75
 Ipomoea batatas  26, 86, 102, 131, 172
 Ipomoea tRicolor  38
 Isoenzymes  100
 Isolation  7, 182
 Isolation techniques  96
 Isoprenoids  5
 Isoquercitrin  146
 Isothiocyanates  173
 Japan  74
 Juglans nigra  218
 Juglone  118
 Juvenile hormones  144
 Karnataka  73
 Kenya  171
 Kernels  32
 Ketones  5, 222, 243
 Keys  75
 Kuwait  230
 Laboratory methods  65
 Laboratory rearing  135
 Lactones  31
 Lactuca sativa  6, 19, 39, 41, 55, 60, 68, 74, 197, 206
 Larvae  9, 78, 79, 84, 90, 100, 110, 128, 133, 135, 144, 167,
 174, 204, 212, 235
 Lasioderma serRicorne  90, 97
 Lawns and turf  199
 Leachates  24, 30, 34, 42, 47, 65, 73, 136, 141, 165, 166,
 189, 206, 230 Leaf age  110
 Leaf area  120, 158
 Leaf water potential  220
 Leaves  13, 14, 21, 24, 30, 32, 35, 43, 64, 65, 105, 109, 111,
 120, 126, 146, 160, 166, 174, 179, 182, 189, 217, 241
 Leguminosae  177
 Lemna minor  39, 118
 Lepidium sativum  4, 21, 168, 170, 231
 Lepidoptera  133, 138
 Lespedeza cuneata  125
 Leucaena leucocephala  31
 Life cycle  147
 Light relations  164
 Limax maximus  226
 Limonoids  235
 Lines  204
 Linum  239
 Linum usitatissimum  239
 Liriodendron tulipifera  13
 Literature reviews  15, 70, 71, 72, 75, 76, 87, 92, 96, 161,
 162, 177, 178, 221, 227, 238
 Litter (plant)  125, 179, 243
 Live mulches  77
 Lobesia botrana  137
 Locusta migratoria  232
 Lolium multiflorum  198
 Lolium perenne  152
 Lycopersicon esculentum  116, 128, 203, 242
 Lymantria dispar  13, 14, 140
 Macrosiphum rosae  185
 Macrotyloma uniflorum  189
 Magnesium  179
 Maize  111
 Mamestra configurata  235
 Manduca sexta  66, 67, 84, 123, 154
 Maturation period  128, 233
 Medicago sativa  29, 47, 89, 95, 152, 175, 198, 216
 Medicarpin  175
 Medicinal plants  19, 171, 215
 Melanoplus sanguinipes  176
 Melilotus indica  166
 Mesembryanthemum crystallinum  179
 Metabolic detoxification  28, 71, 101, 114, 124, 176, 193
 Metabolism  84, 100, 133, 157, 167, 177, 225
 Metabolites  2, 84, 113, 145, 177, 221, 222, 225, 237
 Metamorphosis  69
 Methoprene  144
 Metopolophium  70
 Mexico  127
 Michigan  95
 Microbial activities  3, 52
 Microbial degradation  97, 178
 Microbial pesticides  223, 238
 Microsomes  109
 Midgut  66
 Mineral content  179
 Miscanthus transmorrisonensis  23
 Mississippi  204
 Missouri  107
 Mites  227
 Mitochondria  118, 224
 Mode of action  43
 Molecular conformation  161, 228
 Molting hormones  69, 123, 161
 Monoterpenes  234
 Mortality  90, 147, 154, 194
 Mucilages  170
 Mutants  61
 Mycoherbicides  80
 Mycorrhizal fungi  209
 Mycorrhizas  180
 Mycotoxins  97, 162, 227
 Myrcene  144
 Mythimna separata  10
 Naringin  225
 Natural enemies  227
 Necroses (plant)  6
 Nectar  96
 Nematode control  77, 151
 Nerium oleander  133
 New Jersey  106
 New South Wales  103, 104
 New York  21
 New Zealand  186
 Nicotiana  154, 184
 Nicotine  67, 100, 154
 Nigeria  19
 Night temperature  164
 Nitrate  32, 33, 220
 Nitrate nitrogen  32
 Nitrification  32, 33, 121
 Nitrification inhibitors  32
 Nitrites  32, 121
 Nitrobacteraceae  121
 Nitrogen  27, 33, 91, 159, 183, 217
 Nitrogen content  126
 Nitrogen fertilizers  91, 125, 132
 Nitrogen fixation  32, 215
 Nitrosolobus  121
 Nitrosomonas  121
 No-tillage  16, 156
 Noctuidae  10
 North Carolina  18, 22
 Nuphar lutea  39
 Nutrient content  110, 179, 183
 Nutrient deficiencies  183
 Nutrient solutions  33
 Nutrient uptake  27, 179, 215
 Nutrients  78
 Oats  52
 Ocimene  144
 Ocimum Americanum  37
 Odocoileus hemionus  155
 Ohio  52
 Oilseeds  17
 Oklahoma  88
 Old fields  106
 Organic compounds  33
 Organic sulfur compounds  94
 Oryza sativa  183, 233, 244
 Osmotic pressure  39
 Ostrinia nubilalis  84, 147
 Oviposition  10
 Oviposition attractants  184
 Oviposition deterrents  184
 Oxidation  32, 33, 100, 121
 Oxidoreductases  84
 Oxygen  71
 Oxygenases  100, 123, 167
 P-coumaric acid  121, 239
 Pakistan  194
 Palatability  155
 Panicum antidotale  34
 Panicum coloratum  34
 Panicum miliaceum  91, 102
 Papilio  163
 Papilio polyxenes  71, 212
 Papilionidae  85
 Parasites of insect pests  67, 69, 154
 Parasitic plants  227
 Parasitoids  142
 Parthenium argentatum  42
 Parthenium hysterophorus  31, 73, 227
 Paspalum notatum  125
 Pastureplants  198
 Pastures  49, 152
 Pathogens  227
 Paxillus involutus  27
 Pelargonium  134
 Penicillium  145
 Perennial weeds  75
 Periderm  102, 131, 172
 Peridroma saucia  176, 235
 Peroxidase  185
 Persistence  200
 Pest control  76, 77, 211
 Pest resistance  9, 14, 70, 72, 76, 79, 115, 128, 129, 134,
 150, 183, 192, 204, 217, 233
 Pesticide residues  2
 Pesticide resistance  97, 148
 Pesticides  82, 97, 148
 Phagostimulants  5
 Phalaris aquatica  152
 Pharbitis hederacea  18
 Pharbitis purpurea  102, 109
 Phaseolus lunatus  189
 Phaseolus vulgaris  4, 30, 116
 Phenolic acids  18, 81, 120, 187, 239
 Phenolic compounds  23, 32, 43, 53, 89, 90, 139, 140, 143,
 166, 221, 230
 Phenolic content  40
 Phenols  185, 217
 Pheromones  238
 Phloem loading  201
 Phosphorus  159, 180, 183
 Photosynthesis  118
 Phototoxins  150
 Phylloplane fungi  113
 Physalis  184
 Physicochemical properties  7
 Phytoalexins  225
 Phytophthora  225
 Phytotoxicity  6, 21, 29, 40, 42, 43, 47, 51, 52, 54, 56, 57,
 64, 81, 89, 95, 96, 98, 103, 104, 114, 122, 130, 141, 160,
 168, 175, 188, 189, 191, 206, 214, 216, 219, 222, 225
 Phytotoxins  11, 23, 60, 89, 90, 139, 141, 181, 190, 223
 Picea engelmannii  41, 217
 Pinus banksiana  110
 Pinus contorta  41
 Pinus strobus  110
 Pinus sylvestris  27, 28, 160
 Piper guineense  19
 Pistia stratiotes  196
 Pisum sativum  68
 Plant analysis  119
 Plant communities  23, 87, 107, 230
 Plant competition  45, 46, 87, 106, 132, 177, 186, 207, 226
 Plant composition  3, 4, 5, 7, 11, 12, 13, 14, 17, 20, 38, 43,
 55, 64, 67, 71, 83, 88, 94, 108, 127, 129, 137, 138, 143, 149,
 155, 159, 161, 172, 173, 181, 182, 183, 192, 193, 196, 201,
 203, 211, 222, 225, 229, 231, 233, 241, 243
 Plant density  61, 93
 Plant development  30, 198, 225
 Plant disease control  77, 113
 Plant ecology  106, 226
 Plant extracts  19, 20, 21, 24, 27, 32, 37, 38, 41, 62, 76,
 123, 131, 146, 147, 165, 175, 184, 189, 194, 205, 239
 Plant glands  204
 Plant growth regulators  76, 119, 221
 Plant interaction  45, 61, 96, 205
 Plant metabolism  215
 Plant morphology  70, 216
 Plant nutrition  70, 112
 Plant organs  179
 Plant pathogenic fungi  127, 225
 Plant pests  67, 76, 78, 147, 157, 194
 Plant physiology  58, 165, 221
 Plant pigments  215
 Plant products  161
 Plant protection  80, 82, 157, 215, 236
 Plant residues  29, 42, 60, 64, 125, 132, 178, 219
 Plant secretions  28
 Plant succession  106, 202, 207
 Plantago lanceolata  112
 Plantcomposition  119
 Plantparasitic nematodes  77
 Plants  56, 57, 133
 Plowing  95, 156
 Pluchea  139, 143, 165
 Poa annua  136, 226
 Poa pratensis  95
 Pollen  96, 188
 Pollination  11
 Population density  156
 Population dynamics  151, 156, 233
 Populus  140
 Populus tremula  28
 Potassium  159, 179, 183, 220
 Powders  32, 33
 Predator prey relationships  192
 Predators of insect pests  157
 Predatory arthropods  192
 Proboscidea (martyniaceae)  98
 Proboscidea louisianica  88
 Processing  42
 Products  171
 Propionic acid  228
 Prosopis cineraria  205
 Prosopis juliflora  205
 Protease inhibitors  86
 Protein  232
 Protein content  105
 Protein digestion  66
 Protein synthesis  24
 Proteinases  66
 Proteolysis  24
 Proton pump  109
 Pseudoplusia includens  128
 Pseudotsuga menziesii  217
 Psoralea  195
 Psoralea macrostachya  195
 Pupae  128
 Quantitative analysis  17
 Quercetin  100, 212, 225
 Quercus  105
 Quercus havardii  132
 Quercus petraea  32
 Quercus robur  32, 33
 Quercus rubra  32
 Quinolizidine alkaloids  201
 Quinones  218
 Radicles  21, 51, 63, 189
 Radioactive tracers  134
 Rangelands  45, 132
 Raphanus sativus  122
 Ratios  111, 217
 Regeneration  160
 Research  223
 Resistance  90, 100, 208
 Resistance mechanisms  100, 208
 Resistance to parasites  86
 Respiration  118
 Responses  80, 105, 135
 Reviews  80
 Rhizobium  62
 Rhizoglyphus robini  148
 Rhizomes  166
 Rhizosphere  53, 153
 Rhizosphere fungi  62, 145
 Rhopalosiphum  70
 Rhus  202
 Root analysis  215
 Root exudates  12, 55, 68, 99, 130, 190, 214, 224
 Root hairs  216
 Root meristems  6
 Root nodules  166
 Root shoot ratio  27
 Root systems  216
 Root tips  27
 Root treatment  158
 Roots  6, 16, 27, 30, 35, 37, 40, 41, 47, 74, 111, 131, 168,
 173, 179, 180, 214, 219, 220, 237, 239
 Rorippa  173
 Rorippa sylvestris  55
 Rosa  185
 Rotations  47, 92, 93, 111, 156
 Rotylenchulus reniformis  151
 Rubiaceae  37
 Rudbeckia occidentalis  41
 Rutoside  154
 Saccharomyces cerevisiae  234
 Salicylic acid  124, 237
 Saliva  185
 Sandy soils  159, 230
 Sapindaceae  24
 Saponins  14
 Sasa  53
 Scanning electron microscopy  216
 Schizachyrium scoparium  159
 Schizaphis  70
 Screening  200
 Scrub  159
 Searching behavior  192
 Seasonal fluctuations  217
 Seasonal variation  21, 179, 218, 241
 Secale cereale  2, 3, 4, 99, 222
 Secondary metabolites  112, 133, 162
 Secretion  66
 Seed banks  93, 156, 202
 Seed dispersal  202
 Seed germination  11, 20, 24, 25, 29, 30, 31, 34, 35, 37, 41,
 42, 47, 51, 52, 55, 73, 75, 102, 122, 125, 139, 141, 143, 166,
 170, 172, 175, 189, 191, 195, 203, 205
 Seed predation  202
 Seedbed preparation  132
 Seedgermination  28
 Seedling emergence  60, 89, 104, 125, 152, 175
 Seedling growth  19, 20, 27, 29, 38, 122
 Seedling stage  91
 Seedlings  12, 30, 31, 35, 55, 60, 64, 68, 89, 95, 111, 117,
 120, 126, 132, 160, 169, 174, 186, 198, 205, 214, 220
 Seeds  21, 32, 33, 38, 63, 73, 108, 127, 170, 195, 231
 Selenium  81
 Semiochemicals  169
 Senecio vulgaris  226
 Senses  211
 Separation  229
 Sequential cropping  47
 Sesquiterpenes  83, 182
 Sesquiterpenoid lactones  7, 13, 191, 197
 Setaria (gramineae)  180
 Setaria faberi  156
 Setaria viridis  164
 Sexual reproduction  75
 Shade  164
 Shoots  30, 40, 47, 89, 111, 124, 216, 239
 Sinapis alba  51
 Site factors  230
 Site types  81
 Sitobion  70
 Smell  137
 Sodium  179
 Sodium nitrite  33
 Sogatella furcifera  183, 233
 Soil  60, 166, 206
 Soil amendments  151
 Soil analysis  2, 18, 36, 139, 218, 243
 Soil bacteria  32, 33, 153, 178
 Soil biology  222
 Soil chemistry  126, 187, 221
 Soil fauna  92
 Soil fertility  92
 Soil flora  28, 92
 Soil fungi  153, 166, 178
 Soil organic matter  92
 Soil properties  165
 Soil solarization  77
 Soil structure  92
 Soil treatment  120
 Soil types  104
 Soil water  164
 Solanaceae  6
 Solanine  6
 Solanum incanum  171
 Solanum nigrum  102
 Sorghum  190
 Sorghum almum  34
 Sorghum bicolor  40, 47, 91, 108, 219, 224
 South Carolina  131
 Sowing  95
 Sowing date  103
 Soy straw  52
 Soybeans  52
 Spain  206
 Spatial distribution  179, 226
 Species  81
 Species diversity  107
 Spectral analysis  171
 Spectral data  127, 182
 Spodoptera eridania  71, 163, 212
 Spodoptera exigua  174
 Spodoptera frugiperda  5, 193
 Sprout inhibition  124
 Stemflow  65
 Stems  35, 179
 Stereochemistry  144, 231
 Sterols  1
 Strain differences  62
 Strains  100
 Stress factors  163
 Stress response  183, 233
 Structure  127, 182, 231
 Structure activity relationships  200, 228
 Stubble mulching  103, 104
 Sturnus vulgaris  211
 Substrates  100
 Succession  130
 Sucrose  114
 Superoxide dismutase  212
 Suppression  164
 Surveys  227
 Survival  128, 207, 226
 Susceptibility  204
 Sustainability  80, 92
 Sweden  28
 Sweet potato extract  102
 Sweet potatoes  86
 Symbionts  90, 97
 Symbiosis  215, 221
 Synergism  4
 Synthesis  231
 Tagetes patula  151
 Taiwan  23, 43, 210
 Tamil nadu  40
 Tanacetum vulgare  137
 Tannins  105, 146, 217, 232
 Taro  86
 Tassels  9
 Taxonomy  75
 Temperature  52, 114, 164
 Terpenoids  115, 137, 155, 204, 217, 243
 Testas  32
 Texas  34, 132
 Thiocyanates  17
 Thiophene  84
 Tillage  16, 18, 187
 Tolerance  125, 239
 Toona ciliata  235
 Toxic exudates  40, 79, 209
 Toxic substances  148
 Toxicity  4, 63, 78, 79, 107, 147, 200, 234, 235
 Toxins  176
 Transferases  97, 163
 Transpiration  124
 Triazinoneherbicides  91
 Tribulus terrestris  230
 Trichoderma  145
 Trichomes  134
 Trichoplusia ni  71, 193, 212
 Trichothecenes  11
 Trifolium incarnatum  18
 Trifolium pratense  152
 Trifolium repens  152
 Trifolium subterraneum  152
 Triterpenoids  127
 Triticum aestivum  16, 18, 91, 99, 103, 104, 187, 214
 Triticum durum  25
 Triticumaestivum  98
 Trophic levels  67
 Tropics  210
 Trypsin  66
 Trypsin inhibitors  86
 Tubers  83, 131
 Tunnels  174
 Turgor  220
 Typha latifolia  1
 U.S.S.R.  7
 Undergrowth  43, 218
 Uptake  237
 Uttar pradesh  189
 Vanillic acid  122, 153
 Varietal reactions  111
 Varietal susceptibility  125, 183, 233
 Varietal tolerance  111
 Vegetables  42
 Vegetation  121
 Vegetation management  223
 Vegetative period  179
 Vetch  164
 Vicia  164
 Vicia faba  124, 237
 Vigna mungo  24, 40, 189
 Vigna unguiculata subsp. sesquipedalis  165
 Vigor  132, 167
 Vitis vinifera  137
 Volatile compounds  53, 94, 169, 174, 203, 229, 241, 242, 243
 Volatilization  33
 Washington  226
 Water  42
 Water uptake  117
 Weed biology  75, 93, 156
 Weed competition  186, 230
 Weed control  1, 4, 12, 18, 22, 25, 38, 46, 51, 53, 55, 73,
 75, 77, 80, 99, 114, 131, 139, 143, 152, 156, 164, 172, 173,
 187, 199, 223, 224, 227, 238, 240, 244
 Weedcontrol  109
 Weeds  6, 40, 45, 46, 48, 49, 93, 198
 Weight  9, 111
 Weight gain  128
 Wheat soils  16
 Wildlife management  34
 Wood  21
 X ray diffraction  182
 Xanthosoma sagittifolium  86
 Yams  86
 Yeasts  90, 97
 Yield increases  119
 Yield response functions  103, 104
 Zea mays  9, 52, 91, 95, 111, 122, 144, 156, 164, 169, 174,
 188, 224
 Zea mexicana  188