FILE:  <ENT129.11>                                       Comprehensive Account                     [Several minutes are required to download]                             <Navigate to MAIN MENU>

 

 

                                                             EXPERIMENTAL DESIGN & SAMPLING

 

I.  Luck et al. (1999) emphasized that in order to improve success rates in biological control, an understanding of events in past  successful introduction programs is essential. 

 

A.  Successful cases can be used to test hypotheses about predator/prey interactions, and develop criteria for identifying effective natural enemies. 

 

B.  Van Driesche et al. (1991) reviewd the analytical bases developed in the late 1980's to estimate total losses from parasitism.  Thay stated that, "Because the population of an insect stage typically begins to lose members through death or development to the next stage in the life cycle before the entire recruitment to the stage is completed, at no time are all members of the generation present to be counted.  This idea is analogous to a sink partly filled with water (i.e., the population), into which water is flowing (recruitment) and from which water is draining (death or advancement to the next stage)..."  "To construct a life table, we need to know the total numbers that enter a stage (in this analogy, the total amount of water entering the sink). 

 

C.  What biologists typically measure, however, is the number of animals present per sample unit at points in time (which is analogous to the amount of water in the sink at any given time).  Although it is true that the volume of water present at any time is determined by the moment-to-moment balance of cumulative influx minus cumulative outflow, if these latter quantities are not known, it is not possible to determine total inflow from even the most detailed set of observations on the quantity of water in the tank at fixed moments in time.  What is needed is a continuous record of recruitment for the whole period over which animals enter the stage of interest for the generation.  This can be achieved by measuring recruitment for a series of contiguous intervals spanning the whole period when recruitment occurs (e.g., Van Driesche & Bellows 1988)."

 

D.  Van Driesche et al. (1991) continued, "When the goal is to assess not only how many insects enter a given life stage over the course of a generation, but also to determine how many of that number subsequently become parasitized, the problem is compounded because the basic problem discussed above now applies to two quantities that must be measured; i.e., the total number of hosts recruited and the number that subsequently become parasitized.  The linkages between these value are both dynamic and complex..."  "Although there are some systems in which biology and life history characteristics are such as to produce nondynamic systems not subject to these problems (for example, cases where the sampled stage is in a diapause stage and accumulates without loss as, for example, is approximately the case for gypsy moth eggs, because dead or parasitized eggs remain countable) or systems such as some leafminers in which lost insects continue to be traceable in samples through their remains, the majority of insects do have overlapping recruitment and losses.  For these cases, densities and percentage parasitism values seen in samples do not measure adequately the level of parasitoid effect."

 

                II.  Approaches in the evaluation process include  1.  life table analysis, which is a descriptive method;  2.  stage frequency

                      analysis;  3.  direct measurement of recruitment; 4.  death rate analysis; 5.  experimental manipulations in the field. 

 

                                A.  The primary goal is to determine whether regulation of the host population exists and to identify the agents

                                      responsible for  regulation. 

 

B.  Luck et al. (1999) defined regulation as the biological processes involving natural enemies that suppress prey or host densities below levels that prevail in the absence of natural enemies.  It must be determined whether the populations are regulated, measure the level of regulation and identify the forces involved in regulation.  If the populations are not regulated, if the regulation is intermittent, or if the level of suppression is inadequate, then other options to consider are (1) introduction of additional natural enemies, (2) inoculative or inundative releases, (3) development of plant resistance, (4) change the cultural practices, etc. 

 

C.  There are other key references pertaining to measurement of natural enemy impact (Thompson 1955, Richards et al. 1960, Hafez 1961, Kirtitani & Nakasuji 1967, Manly 1974, 1976, 1977, 1989; Ruesink 1975, Russell 1987, Kolodny-Hirsch 1988, Schneider et al. 1988, Van Driesche 1988, Bellows et al.  1989, Gould et al. 1989, Keating 1989, McGuire & Henry 1989, Van Driesche et al. 1989, 1991a,b, Buonaccorsi, J. P. & J. S. Elkinton 1990, Gould, J. R. 1990a,b, Hazzard et al. 1991).

 

III.  There is probably no single method which can provide conclusive evidence that natural enemies are regulating a population. 

 

A.  Natural enemies are not the only factor involved in many interactions, and the plant can significantly affect the natural enemies' ability to regulate (Flanders 1942, Starks et al. 1972, Price ta al. 1980). 

 

B.  Luck et al. (1999) conclude that no research method if free of technical problems, and management decisions are made with insufficient knowledge.  Therefore research aimed at developing an integrated pest management program is a continuous process in which hypotheses are continually being refined and tested (Way 1973).  Classical biological control and augmentive biological control are important IPM tactics, but they must be pursued and expanded to include situations for which they have not ben emphasized (DeBach 1964, 1974, Ridgway & Vinson 1976, Carl 1982).  Indigenous biological control forms the foundation for pest management and therefore must be utilized if IPM is to become more effective.  Its presence in an agroecosystem can be demonstrated by disrupting it with insecticides (Folsom & Brondy 1930, Woglum et al. 1974, Brown 1951, Pickett & Patterson 1953, Ripper 1956, Bartlett 1968, Smith & van den Bosch 1967, Wood 1971, Ehler et al. 1973, Eveleens et al. 1973, Croft & Brown 1975, Luck & Dahlsten 1975, Luck et al. 1977, Reissig et al. 1982, Kenmore et al. 1984), or by comparing unsprayed, abandoned orchards with treated orchards.  Insecticidal disruption provides one of the best experimental techniques for evaluating natural enemies.  It can reveal the amount of control provided by indigenous entomophages (Stern et al. 1959, Smith & van den Bosch 1967, Falcon et al. 1968, MacPhee & MacLellan 1971, Wood 1971, Flint & van den Bosch 1981, Jones 1982, Metcalf & Luckmann 1982, Kenmore et al. 1984).

 

                               C.  In the experimental evaluation of biological control, testing whether regulation exists and which natural enemies are

                 responsible for the regulation, life tables and their analyses provide a quantitative framework in which to explore the consequences

                of a predator/prey interaction and to generate hypotheses.  However, life tables cannot demonstrate the efficacy of natural enemies

                in suppressing a host or prey population in the field; only experimental methods can do this (Luck et al. 1999).  Some populations

                cannot be manipulated with available technology because they are based on untested assumptions.  Evidence is that natural

                enemies suppress host/prey populations and experimental results suggest that a host plant's nutritional quality, its physical

                structure and its chemical defenses play a role in pest suppression (Denno & McClure 1983, Futuyma & Peterson 1985, Whitham et

               al. 1984, Mattson 1980). 

 

               IV.  The development of an appropriate sampling routine is essential for the evaluation of natural enemies.

 

A.  The design is determined by the objectives of the experiment, the biology of the organisms involved and the cost of acquiring the information to meet the objectives.  The sampling procedure used to acquire data and the statistical techniques used to analyze data must be decided before field evaluation begins. 

 

B.  Appropriate experimental designs require preliminary studies to identify variation sources.  Preliminary samples can save time and resources (Green 1979).  For example Legner (1979, 1983, 1986) and Van Driesche (1983) described some of the problems associated with estimating and interpreting percent parasitism from field samples, while Van Driesche & Bellows (1988) discussed analytical procedures for dealing with some of the problems. 

 

C.  Statistical randomness is important in population sampling and in the assignment of treatments.  Randomness includes locating field plots and selecting sample plants and sample units.  Each sample unit must have an equal chance of being selected.  Nonrandom sampling makes analysis of the data questionable because of the uncertainty associated with the estimation of the values.  Texts and articles on sampling and experimental design should be consulted before an evaluation of natural enemies or of biological control is begun (Morris 1955, 1960, Cochran 1963, Stuart 1976, Elliot 1977, Jessen 1978, Southwood 1978, Green 1979).

 

               V.  Evaluation in biological control must consider the following:  Do natural enemies affect pest population densities; what natural

                     enemies kill a pest; how quickly will an natural enemy kill a pest; how many pests will a natural enemy kill; how does an natural

                     enemy respond to changes in pest densities in the field; and how do environmental changes affect the predator-prey/parasitoid-

                     host interaction (Luck et al. 1999).

 

A.  When evaluating indigenous natural enemy populations, it is necessary to determine whether biological control of the hosts exists. 

 

B.  An effective means compares pest densities in an area not treated with pesticides to pest densities in an area subjected to traditional pesticide practices.  Ceasing the use of pesticides in parts of a field does not constitute a previously unsprayed area, as prolonged pesticide use reduces natural enemies and alternate prey or hosts upon which the natural enemies depend. 

 

C.  Time is required to reestablish interactions between natural enemy and prey/host populations. 

 

D.  Also, the untreated area must be large enough to buffer the plots from pesticide drift and to insulate arthropod populations within from the dynamics and interactions of those in the adjacent areas.  Estimating the degree of regulation exerted by the natural enemies residing in plots subjected to disruptive effects almost always underestimates the amount of potential biological control (Luck et al. 1999).  Pesticide trials in which a small untreated block within a sprayed area is used to estimate the amount of control from factors other than the pesticide treatments are not adequate in that populations in the unsprayed area are overwhelmed by the dynamics of those in the surrounding treated blocks.

 

                VI.  Techniques For Evaluation

 

                               A.  Introduction / Augmentation of Natural Enemies.--In classical and indigenous biological control, a prey population is expected to be self sustaining.  Control derived from augmentive releases is only temporary, lasting one season or less.  The evaluation of each method poses different problems.  In Classical biological control a natural enemy's impact can be demonstrated by comparing the change in a pest's density in the initial release sites with a control site of similar characteristics but lacking the natural enemy (Huffaker et al. 1962, Legner & Silveira-Guido 1983).  A drop in the pest's abundance in the release site compared with the control site suggests that the natural enemy is responsible for the pest's decline.  This conclusion is further supported if the pest's density in the control site also declines following the subsequent introduction or immigration of the natural enemy to that site.  Replication of release and control sites adds confidence to the evaluation if the pattern of decrease is consistent across the experimental plot.  A similar design can evaluate augmentive releases, but the results may be confounded if closely related or morphologically similar indigenous and released natural enemies attack the same pest (see Legner & Brydon 1966).  However, Oatman & Platner (1971, 1978) showed that release and control plots are never identical ecologically.  Exclusion, inclusion or interference methods are required to assess the difference between resident and released natural enemies.  Introducing genetically marked individuals that differ from the resident population only in the genetic marker can also distinguish between resident and introduced populations (Legner et al. 1990, 1991; Luck et al. 1999). 

 

 

The translocation of natural enemies to areas invaded by pest species and subsequent classical biological control gives additional proof that indigenous natural enemies can have a significant role in regulation of native populations (Wilson 1960, Dowden 1962, McGugan & Coppel 1962, McLeod 1962, DeBach 1964, CIBC 1971, Greathead 1971, Laing & Hamai 1971, Rao et al. 1971, Clausen 1978, Luck 1981, Kelleher & Hulme 1984, Cock 1985).  Further proof is given when the introductions are repeated at several locations with similar results (DeBach 1964, Laing & Hamai 1976).

 

 

B.  Exclusion / Inclusion of Natural Enemies.--Cages and other barriers have been used in exclusion and inclusion procedures to evaluate natural enemies (Smith & DeBach 1942, DeBach et al. 1949, DeBach 1955, Sparks et al. 1966, Lingren et al. 1968, Way & Banks 1968, van den Bosch et al. 1969, DeBach & Huffaker 1971, Ashby 1974, Campbell 1978, Richman et al. 1980. Aveling 1981, Faeth & Simberloff 1981, Frazer et al. 1981b, Jones 1982, Elvin et al. 1983, Chambers et al. 1983, Linit & Stephen 1983, Barry et al. 1984, Kring et al. 1985).  Cages to exclude natural enemies were first deployed by Smith & DeBach (1942), using paired sleeve cages to test whether the introduced parasitoid Metaphycus helvolus (Compere) regulated the black scale, Saissetia oleae (Bern.).  Comparison of the black scale in the open and closed cages showed that less black scale survived in the open cages.  This technique was modified by using insecticide impregnated netting to kill natural enemies that emerged in the closed cages when the methods was used to evaluate other classical biological control projects (DeBach et al. 1949, DeBach 1955, DeBach & Huffaker 1971). 

 

  1.  Cages with different sizes of mesh have been used to exclude natural enemies based on their size (Campbell 1978, Kring et al. 1985).  Three types employed were (1) a complete exclusion cage with small mesh netting and sealed at both ends, (2) a control cage with similar netting and open at both ends and (3) a partial exclusion cage with large mesh netting and closed at both ends.  The latter excluded large predators but allowed access of small predators and parasitoids. 

 

 

                  2.  Sleeve and field cages with more complex designs, such as those which enclosed whole plants, accompanied by samples of the prey and natural enemy populations, showed that the spring increase of predators eliminated black bean aphid, Aphis fabae Scop., colonies on its overwinter host, Euonymus europaenus L., after June (Way & Banks 1968).  If spring aphid populations had been dense on the tree, the predators that remained after the aphids emigrated to their summer hosts prevented recolonization of spindle tree by late fundatrices during the summer, even though the spindle tree was capable of supporting an increasing aphid population.  Closed field cages covered with dieldrin treated netting coupled with hand removal excluded natural enemies from some spindle trees whereas open field cages constructed with slatted walls allowed access of the natural enemies to the aphids on the uncaged trees but provided the same degree of shading as the closed cage (Way & Banks 1958, 1968).  Such experiments and making census of populations on the sample twigs document the importance of predators in excluding aphids from the overwintering host plant during the summertime (Luck et al. 1999).

 

 

                 3.  The evaluation of indigenous natural enemies of cereal aphids was done in large field cages and accompanying population samples.  The experimental design combined field cages erected at several intervals after the aphids immigrated into a winter wheat field.  The growth rates and peak densities of the aphid populations within the cages was compared with those in several open plots of similar size (Chambers et al. 1983).  Samples showed that the abundance of Coccinella 7-punctata L. was negatively correlated with aphid abundance in the open plots but the incidence of parasitism and disease was not negatively correlated with aphid abundance.  These latter two factors were more common in the caged plots.  If the difference between the aphid densities in the cage and open plots was converted to per capita aphid consumption, based on the sampled coccinellid densities, the calculated values were within the range of known values.  Coccinellids appeared to be the key agents limiting the growth rate and peak abundance of cereal aphids during mid season but they were unable to do so early in the season (Rabbinge et al. 1979, Carter et al. 1980).

 

                  4.  Field cages with open field controls were used to determine whether the predator complex aggregated at dense patches of the pea aphid, Acrythosiphon pisum (Harris) (Frazer et al. 1981b).  The cages excluded the predators and allowed the aphid population to increase to about 5X that of the open control plots.  When the cages were removed the aphid populations declined to the densities that prevailed in the control plots and the decline was correlated with increased predator numbers aggregating at the denser aphid patches.  Large field cages have also been used to evaluate the potential of predators in cotton to reduce egg and larval populations of the tobacco budworm, Heliothis virescens (F.) (Lingren et al. 1968).  Evening releases of budworm moths initiated the prey populations within the cages.  Fewer prey survived in the cages with predators than in cages excluding predators.  Similar studies were conducted in California cotton to evaluate predation on the survival of larval populations of the cotton bollworm, Heliothis zea (Boddie) (van den Bosch et al. 1969).  The cotton plants within the predator-free cages were treated with an insecticide to eliminate resident predators before bollworm larvae were introduced.  Significantly fewer prey survived in the untreated cages and significantly more predators were collected from the untreated cages.

 

  5.  In order to determine whether indigenous natural enemies or microclimatic changes within a cage explained the increased survival of caged European corn borer, Ostrinia nubialis (Hübner) larvae, caged and uncaged plots and plots of similar size but enclosed with a cage within a cage were used (Sparks et al. 1966).  The double cages was designed so that the screened panels on the inside cage were opposite that unscreened panels on the outside cage and vice versa.  This arrangement allowed predators access to the plants inside while maintaining the same level of shading and air flow in both the complete cage and cage within a cage plots.  Entomopathogenic fungi (Deuteromycotina) effects were also tested with cages for the black bug, Scotinophara coarctata F., in rice (Rombach et al. 1986a.).  Adult bugs were introduced into screened cages and applications of fungi Beauveria bassiana (Bals.) Vuill, Metarhizium anisopliae (Metsch.) and Paecilomyces lilacinus Thom. were made with a backpack sprayer.  The black bugs were significantly less abundant in all treatments when compared with untreated controls, with effects lasting to nine weeks.  Similarly caged brown planthoppers, Nilaparvata lugens Stal, were treated with entomopathogenic hyphomycetes (Fombach et al. 1986b).  Mortality from fungal infections ranged from 63-98% three weeks after application. 

 

  6.  Ground predators, principally carabids, were excluded with trenches that contained insecticide soaked straw, from the cabbage root fly, Erioischia brassicae (Bouché) (Wright et al. 1960, Coaker 1965).  Polythene barriers were used to exclude predators from two of three treatments in which the predator density was manipulated to determine its effect on the density of aphid populations (Winder 1990).  Sticky bands around selected branches of a spindle tree were used to exclude the walking predators of Aphis fabae (Way & Banks 1968) and around the plant base to exclude walking predators of Trichoplusia ni (Hübner) (Jones  1982).  Sticky circles around Trichoplusia ni eggs were used to exclude predators and parasitoids from attacking the eggs (Jones 1982).

 

  7.  Studies relating cage densities to the densities of resident field populations of predators outside the cages have been used for aphids and Lepidoptera (Frazer & Gilbert 1976, Campbell 1978, Aveling 1981, Frazer et al. 1981b, Chambers et al. 1983), providing useful hypotheses (Way & Banks 1968, van den Bosch et al. 1969, Campbell 1978, Carter et al. 1980, Aveling 1981, Faeth & Simberloff 1981, Frazer et al. 1981b, Chambers et al. 1983).  Cages can provide quantitative information on predation rates (Elvin et al. 1983) but not without limitations.  Small sleeve cages inhibit predator or prey movement and are good for experiments with sessile species or species with low vagility (smith & DeBach 1942).  The abundance of citrus red mite, Panonychus (= Metatetranychus) citri (McG.), within sleeve cages was sometimes 12X greater than outside sleeve cages (Fleschner 1958) even though the mite population outside the cages was kept predator-free by continuous hand removal of predators.  It was thought that the cage prevented the reproductive females from emigrating, that the microclimate within the cages favored rapid growth of the mite population, or both factors influenced population growth (Fleschner et al. 1955, Fleschner 1958).

 

  8.  It is not possible to identify which members of a predator/parasitoid complex are regulating a host population with exclusion cages unless the complex consists of one or a few species (Jones 1962).  Partial exclusion cages may show whether small predators, pathogens or parasitoids regulate in the absence of large predators, but they cannot show whether large predators regulate prey in the absence of parasitoids or small predators (Luck et al. 1999).  Cages may also inhibit predator or prey movement or interfere with natural enemy oviposition.  Two leaf mining species on oak failed to reproduce within whole tree cages and a third species failed to reproduce in one cage (Faeth & Simberloff 1981).  Aphid alates cannot emigrate from a cage, thus caged versus uncaged aphid populations may show differences in density because alate immigration reduces the uncaged aphid population.  Some predator species aggregate at patches of high prey density in a numerical response (Readshaw 1973, Frazer et al. 1981a. Kareiva 1985).  Such behavior may be inhibited by cage size because the spatial pattern in nature to which the predator species responds is larger than that present within the cage.  Also, confining predators to a cage may causae them to search areas more frequently and thereby increases the likelihood that they will encounter prey.  Under these conditions the predator may reduce prey densities to levels below normal, and in this way inclusion studies resemble laboratory experiments in which predators are confined with prey (van Lenteren & Bakker 1976, Luck et al. 1979). 

 

  9.  Erroneous interpretations can result when prey are placed into a cage without consideration of their preferences for oviposition sites, their density and distribution patterns or their preferred feeding sites under field conditions.  Some predators and parasitoids use kairomones to find their prey and hosts (Hassell 1980, Nordlund et al. 1981).  Some kairomones are associated with feeding activity.  Placing prey or hosts in new sites influences their risk of detection.  Food quality may affect a phytophage's feeding time and increase its risk to predation because of the kairomones released while feeding (Nordlund et al. 1981).  Detailed studies of a predator's searching behavior and capture rates and a prey's oviposition and feeding behavior are important (Fleschner 1950, Dixon 1959, Frazer & Gilbert 1976. Gilbert et al. 1976, Rabbinge et al. 1979, Carter et al. 1980. Baumbaertner et al. 1981, Frazer & Gill 1981, Sabelis 1981).

 

  10.  Whenever predator free controls are employed, it is difficult to exclude all predators, even when they have been treated with insecticides (van den Bosch et al. 1969, Irwin et al 1974, Elvin et al. 1983).  Some predators may pass through excluding screens when in small developmental stages (Sailer 1966, Way & Banks 1968), or they are difficult to exclude because they become buried in the soil (Frazer et al. 1981a, Elvin et al. 1983).  Cages also alter the microclimate through shading and inhibiting air flow.  Exclusion and partial exclusion cages using terylene netting reduced the light intensity inside cages by 24-37% (Campbell 1978) and saran screen reduced solar radiation by 19% (Hand & Keaster 1967).  Such shading required the use of a more shade tolerant cotton cultivar than was normally planted in the region (van den Bosch et al. 1969).  Shading also affects plant physiology and thus may affect the plant's quality as a substrate for the host or prey population (Scriber & Slansky 1981).  Temperatures within cages used in a corn borer study were 8-10°F lower than the temperature outside.  The humidity fluctuated more moderately within and was 5-10% higher than that outside the exclusion cages (Sparks et al. 1966). 

 

 

  11.  Solar radiation changes cause differences in leaf temperature by as much as 13°C (Hand & Keaster 1967).  Leaf temperatures and moisture availability influence photosynthetic rates and evapotranspiration (Gates 1980).  Leaf temperatures probably affect the behavior and feeding rates of phytophagous hosts and prey.  Temperature related interactions between the growth rates of aphids and the searching rates of their predators are important (Frazer & Gilbert 1976, Frazer et al. 1981a).  Screening also reduced wind speed within a cage by as much as 48% (Hand & Keaster 1967) which, depending on RH and wind velocity outside and inside a cage, influences the leaf's boundary layer within the cage (Gates 1980, Ferro & Southwick 1984).  Instrumentation allows the monitoring of many of these effects but their influence on predator/prey interactions must be assessed (Luck et al. 1999).

 

C.  Removal by Insecticide Treatment.

 

  1.  Natural enemy complex impact may be assessed through the application of insecticides.  The method was first used to kill natural enemies of the long-tailed mealybug, Pseudococcus longispinus (Targ.), without affecting the mealybugs (DeBach 1946).  Insecticides have been used to determine whether indigenous predator populations in cotton suppress populations of the beet armyworm, Spodoptera exigua (Hübner), and cabbage looper, Trichoplusia ni (Ehler et al. 1973, Eveleens et al. 1973).  Early season insecticides applied to cotton were thought to interfere with natural controls (Ehler et al. 1973, Eveleens et al. 1973).  Large blocks (3-4 square miles) were treated with an insecticide scheduled during early season, early and midseason and early, mid- and late season.  A fourth plot served as an unsprayed control.  Samples and observations showed that the absence of predators in the treated plots was correlated with the increased survival of beet armyworm eggs and first generation small larvae of the cabbage looper.  The hemipteran predators, Geocorus pallens Stal, Orius tristicolor (White) and Nabis americoferus Carayon were implicated as the most important predators since they were the most affected by the treatments whereas Chrysoperla carnea Stephen was not so strongly affected.  Insecticide treatment showed that the suppression of cabbage looper densities in celery arising from egg parasitism by Trichogramma spp. and predation of eggs and young larvae by Hypodamia convergens Guer. and O. tristicolor was sufficient to prevent economic damage before the production of the first marketable petiole in celery (Jones 1982).

 

  2.  Insecticides were also used to test whether the coccinellid, Stethorus sp. regulated the density of the two spotted spider mite, Tetranychus urticae (Koch), in a previously untreated apple orchard in Australia (Readshaw 1973).  Two applications of malathion increased the density of the mite populations.  Tetranychus urticae, unlike the predator fauna associated with it, was resistant to malathion.  Stethorus regulated the mite population by numerically responding both aggregatively and reproductively to the denser mite patches.  Even with insecticide disruption and stimulation of the mite reproduction (Chaboussou 1965, Bartlett 1968, van de Vrie et al. 1972, Dittrich et al. 1974), Stethorus was able to prevent the mite population from attaining an economic density of 100 mites/leaf on most trees.

 

 

  3.  The action of two parasitoids of the olive scale, Parlatoria oleae (Colvee), was evaluated using insecticides (Huffaker & Kennett 1966).  This scale is bivoltine on olive in the San Joaquin Valley of Calviornia.  One generation occurs during the autumn and spring and the second generation during summer.  Aphytis paramaculicornus DeBach & Rosen and Coccophagoides utilis Doutt was introduced for biological control (Rosen & DeBach 1978).  Aphytis dominated during the autumn and spring scale generation whereas Coccophagoides dominated during summer.  Three DDT treatments were used to exclude the parasitoids: (1) a spring treatment to exclude Aphytis, (2) as summer treatment to exclude Coccophagoides and (3) a spring and summer treatment to exclude both parasitoids.  Untreated trees were left as controls.  It was thought that DDT residues on the foliage and twigs inhibited the parasitoids but did not affect the scale's reproduction and survival.  Treatments which excluded Coccophagoides had higher scale densities than the untreated controls but lower densities than the treatments which excluded Aphytis.  Treatments that excluded only one of the parasitoids had lower scale densities than treatments that excluded both parasitoids.  Treatments also indicated that together the parasitoids provided better biological control than either did alone even though the mortality contributed by Coccophagoides was only about 5%. 

 

  4.  Inoculation of fumigated (12 hrs with methyl bromide) and unfumigated poultry manure with  Musca domestica L. eggs demonstrated 53.4 to 99.4% mortality in the presence of predatory and scavenger arthropods (Legner 1971).  Significant negative correlations of parasitization with increasing host densities were explained by parasitoid behavior.  Inherently, single female parasitoids without interference from other individuals of the same or different species respond positively with increases in host density; parasitization rates increase, which appears to be correlated with increases in the production of progeny (Legner 1967).  However, when groups of parasitoids concentrate their search among several host pupae, as is common in nature, their efficiency per female is decreased through mutual interference, that apparently involves combinations of physical interruption and chemical effects.  There was some evidence that female parasitoids were strongly attracted to denser concentrations of their hosts in their habitat (e.g., Legner 1969), which evidence further tends toward increases in the interference factor at natural high host densities.  Furthermore, any interference that would deter some female parasitoids from oviposition during the first few days of adult life would lower fecundity and longevity (Legner & Gerling 1967).  Operating collectively, these several forces would tend to produce the observed apparent negative correlation between parasitization and host density. 

 

  5.  Several problems are associated with interpreting results from an insecticide treatment, however.  The pesticide may stimulate reproduction of the prey population.  There may be a pesticide induced sex ratio bias, and pesticide induced physiological effects on the plant may arise.  Mites that are exposed to sublethal doses of some pesticides are stimulated reproductively and occasionally even increase female biased sex ratios (Charboussou 1965, Bartlett 1968, van de Vrie et al. 1972, Dittrich et al. 1974, Maggi & Leigh 1983, Jones & Parrella 1984).  Such effects may also extend to aphids (Bartlett 1968, Mueke et al. 1978), and delphacids (Chelliah et al. 1980, Reissig et al. 1982).  Differential mortality resulting from pesticide treatments has also been reported.  Male black pineleaf scale, Nucalaspis californica (Coleman) (Edmunds & Alstad 1985), and California red scale, Aonidiella aurantii (Maskell) (Shaw et al. 1973) are more susceptible to pesticides than females.  Plant physiology is also affected by insecticide applications (Kinzer et al. 1977, Jones et al. 1983).  Row crops treated with certain insecticides become attractive oviposition sites for Lepidoptera (Kinzer et al. 1977).  Interactions between aphid reproduction, insecticides and cultivars have been reported on alfalfa (Mueke et al. 1978).  Knowledge of the biology and interactions is required to properly time an insecticide application to disrupt the natural enemy populations while minimizing their effects on prey or host.  Because insecticides potentially stimulate arthropod reproduction and effect plant physiology, estimates of predation rates with this exclusion method should be done cautiously.  Although insecticide treatments stimulated the brown planthopper, Nilaparvata lugens Stal, reproduction, the amount of stimulation could not account for the high levels of resurgence.  Only the reduction of natural enemies could.  Insecticides can be used to determine the relative importance of natural enemies when the complex is composed of a few species showing temporal separation of their effects, in seasonal occurrence or in the generations they attack (Luck et al. 1999).

 

 

D.  Removal of Natural Enemies by Hand.

 

  1.  Although laborious, hand removal has been used to evaluate the predators of tetranychid mites on citrus and avocado and to compare results obtained with other exclusion methods (Fleschner et al. 1955, Fleschner 1958).  It has also been used to evaluate the mirid, Crytorhinus fulvus Knight, introduced to control the taro leafhopper, Tarophagus proserpina (Kirkaldy) (Matsumoto & Nishida 1966).  Predation of Aphis fabae was also assessed in part by removing adult predators by hand when they flew onto predator free branches (Way & Banks 1968).  A sticky band at the base excluded walking predators from feeding on A. fabae individuals placed on the branch.

 

  2.  Luck et al. (1999) believe that the hand removal method deserves more attention, especially as a method of checking for bias in other exclusion methods.  However, it seems to be limited to studies of predator/prey interactions with species of low vagility, those that occur at reasonable densities and are diurnally active or are undisturbed by night lights (Luck et al. 1999).

 

E.  Prey Enhancement.

 

 

  1.  Prey may be placed directly on plants in the field to stimulate predator attraction.  This procedure involves tethering prey to a substrate (Weseloh 1974, 1982) or placing them on leaves or other plant parts where they would normally occur (Ryan & Medley 1970, Elvin et al. 1973, van Sickle & Weseloh 1974, Weseloh 1974, 1978, 1982; Torgensen & Ryan 1981).  Some studies marked the prey with dyes before placing them in the field (Hawkes 1972, Elvin et al. 1973).  The prey were visited frequently to measure predation, and if predation was observed, the predator's identity was noted.  Predators such as spiders can be observed in the field with their prey *Kiritani et al. 1972), and web spinning spiders leave cadavers in or beneath their webs (Turnbull 1964).

 

  2.  It is sometimes more practical to use greenhouse grown plants of the same age, size and variety as plants used in field studies.  Plants can be caged in the greenhouse or field for pest oviposition.  Then the infested plants are transferred to the field and monitored for parasitism and predation.  Van der Berg et al. (1988) used eggs of several foliage-feeding rice pests to determine predation.  The egg chorion showed that eggs were attacked by predators with chewing or sucking mouthparts.

 

  3.  Predation and parasitism was thought to alternate as principal mortality factors during the year in studies that followed the seasonal incidence of predation and parasitism of eggs of the yellow stemborer of rice, Scirpophaga incertulas (Walker) (Shepard & Arida 1986).  The technique of prey enhancement may be used to advantage with cages and or insecticides.  However, a major limitation is that prey must be limited to sessile forms such as eggs, pupae or some scale insects, although there are possibilities with tethered hosts (Weseloh 1974).  Kairomones and other chemical cues may be important to establishing the appropriate interaction (Nordlund et al. 1981). 

 

 

VII.  Methods For Detecting Predation/Parasitism

 

A.  Serology.

 

                 1.  Predators have been associated with their prey with serological methods (Dempster et al. 1959, Dempster 1960, 1964, 1967; Rothshild 1966, 1970, 1971; Frank 1967, Ashby 1974, Vickermann & Sunderland 1975, Boreham & Ohiagu 1978, Sunderland & Sutton 1980, Gardner et al. 1981, Greenstone 1983).  Predations rates have also been estimated with serology (Dempster et al. 1959, Dempster 1960, 1964, 1967).  A precipitin assay has been also used (Boreham & Ohiagu 1978, Ohiagu & Boreham 1978, Southwood 1978).  Other methods are the enzyme-linked immunosorbent assay (ELISA) (Vickermann & Sunderland 1975, Fichter & Stephen 1979, 1981, 1984; Ragsdale et al. 1981, Crook & Sunderland 1984, Sunderland et al. 1987, Sopp & Sunderland 1989), and an assay based on passive hemagglutination inhibition (PHI) (Greenstone 1977, 1979).  Agglutination assay employs polystyrene latex particles coated with antibody (Boreham & Ohiagu 1978, Ohiagu & Boreham 1978).  Such methods detect prey particles in the gut of predators by its reaction with antibodies obtained from a vertebrate, such as a rabbit, that has been sensitized to the prey.  The reaction is a visible precipitate.  (Also see Boreham & Ohiagu 1978, Miller 1978 and Sunderland 1988).

 

  2.  Detection of prey in a predator's gut is influenced by the size of prey, size of meal, time since the meal was taken, the rate of digestion, whether the natural enemy is a sucking or chewing predator, the abundance of taxonomically closely related prey and the sensitivity of the test.  Sensitivity of the assay can be increased if the antibody is linked to an enzyme (ELISA).  When the antibody reacts with prey, the enzyme carried with the antibody allows amplification of the reaction because one enzyme molecule can convert many molecules of substrate.  This assay may detect hemolymph dilutions of more than 260,000 (Fichter & Stephen 1981) and is often sufficient to differentiate among prey stages (Ragsdale et al. 1981).

 

  3.  Both precipitation and ELISA techniques are useful for identifying the prey in a predator's diet and estimating predation rates (Sunderland 1988).  ELISA is more sensitive to the presence of small amounts of antigen (prey protein or carbohydrate), is suitable for large scale testing and can be used with a minimum of equipment.  Material necessary for the tests may be prepared and stored under refrigeration for six months (Sunderland 1988).

 

  4.  The passive haemagglutination assay (PHA) is a method for increasing sensitivity of the precipitin test.  Sheep red blood cells (rbc) are chemically coated with the antigen of the suspected prey.  Antigen coated rbc's are added to a solution of specific antibody and combine with the antigen molecules on the rbc to form a mat (agglutination).  Small amounts of antibody cause agglutination.  In antibody-free controls the rbc's do not agglutinate and this inhibition forms the basis of the assay.  The amount of antibody required to cause agglutination is determined and added to an extract of a predator's gut contents.  If prey protein or carbohydrate (antigen) is present it binds with the antibody.  When antigen coated rbc's are added, they will not agglutinate because the antigen from the predator's gut has been bound by the antibodies (Luck et al. 1999).  A small amount of antigen produces inhibition which explains the assay's greater sensitivity than that of a comparable precipitin assay (Greenstone 1979).  Freshly sensitized erythrocytes have to be prepared each time an assay is conducted (Boreham & Ohiagu 1978), and this requires skilled operators.

 

                                  5.  The precipitin test was originally used to document arthropod predation of mosquito larvae (Bull & King 1923, Hall et al. 1953, Downe & West 1954) and latter was applied to terrestrial predator/prey interactions (Downe & West 1954).  The first prey for which estimates were attempted from field samples was a chrysomelid beetle Gonioctena (= Phytodecta) olivacea (Forster) feeding on broom (Dempster 1960).  Tests revealed six mirids, two anthocorids, a nabid, a dermaptern and red mites feeding on the beetle in the field.  Laboratory tests showed that only the older mirid and anthocorid stages fed exclusively on younger stages of G. olivacea.  A single laboratory feeding by the mirids and anthocorids could be detected 24 hrs after they had ingested a meal, and feeding by a dermapteran could be detected 60 hrs after it had fed (see Luck et al. 1999).

 

                                  6.  The degree of overlap between older stages of the predator and younger stages of the beetle influenced the number of beetles preyed upon.  Densities of prey and predators were estimated from field samples.  The fraction of positive responses in predator samples estimated the fraction of the predator population that had fed on G. olivacea.  Because G. olivacea were scarce in the field while alternative prey were abundant, encounters between G. olivacea and the predators were infrequent.  Therefore, if a predator tested positive to G. olivacea antibody, it was interpreted as a single predation event.  Then the number of beetles preyed upon by each predator could be estimated suing the equation:

 

                            Pa = (NpiFpiTpi) / Rpi

 

 

where Pa is the number of prey killed; Npi the density of the predator (or stage of predator) i; Fpi the fraction of positive tests of the ith predator in a sample; Tpi the duration in days that the appropriate prey and predator stages are coincident in the field; and Rpi the retention time of a single prey feeding by ith predator (or stage of predator).  Estimates from the precipitin test of egg and larval mortality due to predator for two beetle generations were found to agree closely with the independent estimates of "unknown" losses of eggs and young larvae during the same two beetle generations (Richards and Waloff 1961). 

 

                7.  The precipitin test also was used to identify the predator species and to determine the fraction of Pieris rapae (L.) eggs and young larvae that died due to predation (Dempster 1967).  Because of the relative scarcity of P. rapae a positive precipitin test was interpreted as one predation event.  Studies of the delphacid Conomelus anceps (Germar) employed precipitin tests to identify ten of 91 potential predators (Rothchild 1966).  The precipitin test could not be used to estimate predation rates because multiple predation events were possible. 

 

                 8.  For estimating predation rates with the precipitin test it is necessary to have information about predator and prey densities, densities of alternate prey, the period during which a meal can be detected in each predator and prey and predator stages involved.  Precipitin tests estimate predation rates of prey which form a small fraction of the available prey or infrequent predation events.  A slight bias may arise in such estimates if predators have fed on other predators that have fed on the prey, if a suspected predator is phytophagous but ingests sessile prey stages while feeding on the plant or if a suspected predator feeds on prey carrion (Boreham & Ohiagu 1978).  The precipitin test may also yield biased estimates of predation rates from cross reactions between the antibodies of closely related species.  Therefore, a knowledge of the local fauna which might serve as prey and the predator's propensity for local movement is essential to the successful application of this test (Luck et al. 1999).  Also the serum developed from one prey stage may not react with the antigen of another (Boreham & Ohiagu 1978).  Sufficient resources must be committed in order to use this technique:  prey must be collected in sufficient numbers to elicit an immunological response when injected into the vertebrate.  As such the procedure is not ideal when applied to small prey such as mites (Murray & Solomon 1978).

 

                                9.  When used in conjunction with other population studies, precipitin assays may be very helpful.  Few other methods can provide quantitative estimates of predation rates under natural field conditions.  Although they cannot be used to estimate predation rates under all situations, they are valuable for identifying predator species or stages that feed on a prey.  This method deserves more attention especially as more sensitive tests such as ELISA are available (Vickermann & Sunderland 1975, Fichter & Stephen 1981, 1984; Ragsdale et al. 1981, Crook & Sunderland 1984, Sunderland et al. 1987, Soop & Sunderland 1989).  A great advantage is that predation is allowed to occur naturally.

 

                  B.  Electrophoresis & Isoelectric Focusing.

 

                                1.  Predators may be associated with the prey with electrophoretic techniques.   Electrophoresis separates proteins based on charge and size differences in an electrical field.  Differences in charge and size commonly occur among isoenzymes (proteins catalyzing the same reaction) from different taxa.  If the prey and predator have isoenzymes with different electrophoretic mobilities, the analysis of homogenates prepared from predators fed on prey should exhibit protein bands corresponding to the predator and the prey.  Also if there are several potential prey of a predator, and if the prey have electrophoretically distinct isoenzymes, analysis of predator homogenates can reveal the prey species inside the predator.

 

                                 2.  Electrophoresis can be successful if the prey isoenzymes are detectable after predator feeding, and electrophoretic variation occurs among the prey and predator isoenzymes.  Isoenzyme detection depends on prey size, in vitro activity of the isoenzyme, presence and volume of the predator foregut, and the type of electrophoresis employed (Murray & Solomon 1978, Giller 1984, Lister et al. 1987, Soop & Sunderland 1989).  Electrophoretic variation depends on the suite of isoenzymes available for comparison and the type of electrophoresis.  Standard electrophoretic procedures (starch gel and polyacrylamide gel electrophoresis) can detect prey isoenzyme activity for several isoenzyme types involving relatively large prey (>2-3 mm body length).  Under this size, the number of detectable prey isoenzymes is diminished and hence the chance of distinguishing closely related prey is decreased.

 

                                 3.  Enhanced sensitivity of electrophoretic methods include conventional electrophoresis in cellulose acetate membranes (Easteal & Boussy 1987, Höller & Braune 1988) and isoelectric focusing (IEF) (see Luck et al. 1999).  IEF has advantages over other techniques involving small and large prey.  In IEF, proteins are "focused" into narrow bands along relatively broad pH gradients.  Focusing enhances the detection of enzymes compared to other techniques which gradually spread the proteins into diffuse bands.  In addition, because relatively broad pH gradients are used in IEF, enzymes with different charges, such as may occur between unrelated prey taxa, will remain sharply focused on the gel.  The fine resolution of IEF does not affect the ability to distinguish enzymes with very similar charges.  With standard techniques, these contrasting problems are difficult to solve simultaneously as one set of conditions (buffer type and pH, gel type) may be optimal for one prey type but not others.

 

                                4.  The prey of several arthropod predators have identified with electrophoresis.  Polyacrylamide gradient gel electrophoresis was used to detect prey protein (esterases) in the gut of predators after they had fed on known prey (Murray & Solomon 1978).  The technique detected esterases of Panonychus ulmi (Koch) in the predaceous mite Typhlodromus pyri (Scheuten), and in two anthocorids, Anthocoris nemoralis (F.) and Orius minutus (L.) that had fed on the mite in the laboratory.  Dicke & DeJong (1986) used methods to determine whether T. pyri and Amblyseius finlandicus (Oudemans) also fed on the apple rust mite, Aculus schlechtendali (Nalepa) as an alternate host in the field.  Electrophoresis was also used to identify the prey species exploited by A. nemoralis on alders in the field (an aphid Pterocallis alni [DeGeer]) (Murray & Solomon 1978).  Electrophoresis with polyacrylamide disc gels detected esterases of several prey species in the gut of the waterboatman Notonecta glauca L. (Giller 1982, 1984, 1986).  A meal was detectable from 17-48 hrs depending on temperature and meal size, and was strongly correlated with the length of time the meal spent in the foregut (Giller 1984).  Giller (1986) used electrophoresis to identify the prey of N. glauca and N. viridis Delcourt in the field.  Lister et al. (1987) used polyacrylamide gel electrophoresis and electrophoresis and esterase allozymes to determine the diet of some microarthropods and the predation rate by the acarine predator Gamasellus racovitzai (Trousessart).

 

  5.  Predation rate estimates with serological methods and electrophoresis requires substantial resources.  The techniques call for the development of antibodies or methods for identifying the isozymes of the prey species or stage, the development of methods to estimate the predator and prey densities, including those needed to estimate the densities of alternate prey, and the identification of the predator and prey stages involved.  Initially the use of these techniques to estimate predation rates appeared limited to prey populations which form a small fraction of the available prey or in which predation events by a predator are frequent.  Frequent predation confounds interpretation of a positive test because a single large meal cannot be distinguished from several small meals.

 

                  6.  Immature parasitoids within aphids have been detected with electrophoresis (Wool et al. 1978, Castanera et al. 1983), and in whiteflies (Wool et al. 1984).  The parasitoid Aphidius matricariae Hal. was detected in the green peach aphid Myzus persicae (Sulz) and parasitism of the white fly, Bermesia tabaci (Gennadius) by the endoparasitoids Encarsia lutea (Masi) and Eretmocerus mundus (Mercet), was detected with electrophoresis and histochemical staining for esterases.  But the whitefly parasitoids could not be identified to species (Wool et al. 1984).  Electrophoresis allows the processing of large numbers of hosts to estimate the fraction that are parasitized and sometimes the parasitoid species involved.  This contrasts with the traditional methods in which field samples are dissected while fresh or reared.  Electrophoresis can detect within a host immature parasitoids without dissection and parasitoid enzyme activity within a prey cannot be confused with host's enzyme activity.

 

C.  Marking Prey.

 

                   1.  Predator species and/or predation rates have been identified with marking techniques.  Markers have included radioactive isotopes -151europium (Ito et al. 1972), 32phosphorus (Jenkins & Hassett 1950, Pendleton & Grundmann 1954, Jenking 1963, McDaniel & Sterling 1979, McCarty et al. 1980, Elvin et al. 1983) and 137cesium (Moulder & Reichle 1972), 14carbon (Frank 1967), rare elements (Stimmann 1974, Shepard & Waddill 1976), and dyes (Hawks 1972, Elvin et al. 1983).  Prey are fed (Elvin et al. 1983, Frank 1967, Room 1987) or injected (McDaniel & Sterling 1979, McCarty et al. 1980) with the radioactive isotope and the radioactivity is detected in a predator with scintillation, a Geiger counter, or autoradiography.  For autoradiography suspected predators are collected after exposure to labelled prey and are glued to paper, which is placed against X-ray film (McDaniel & Sterling 1979).  The film is developed, and dark spots on the film produced by the rays from 32phosphorus indicate labelled predators.  Methods involving isotopes require training and necessary equipment to perform the assays.  Safety regulations and environmental considerations may limit the use of the method in some situations.  Other disadvantages, as with electrophoresis and serological techniques, include the inability to detect whether a predator had fed on other predators that had consumed labelled prey or whether a prey was scavenged (Luck et al. 1999).  Experiments using isotopes, especially those using autoradiography, are simpler to conduct than serological and related techniques.  Methods using labelled elements require several manipulations, but they provide more information per unit effort than other kinds of marker tests.

 

                   2.  Such rare elements as rubidium and strontium also have application as labels.  They can be sprayed on foliage or placed in the diet of the prey, incorporated into the prey's tissues and then transferred to the predators or parasitoids who feed on labelled hosts (Stimmann 1974, Shepard & Waddill 1976).  The mark should be retained for life, and self-marking is possible via a labelled plant.  However, the technique requires an atomic absorption spectrophotometer, which is expensive, and placement of the labelled prey on plants may expose them to abnormal predation rates.  Phytophages seldom choose feeding or oviposition sites on their plant hosts at random (Ives 1978, Wolfson 1980, Denno & McClure 1983, Guerin & Stadler 1984, Whitham et al. 1984, Myers 1985, Papaj & Rausher 1987).  Parasitoids and predators do not search their habitats uniformly (Weseloh 1974, 1982; Fleschner 1950).  Therefore, without the proper behavior studies, the degree of bias in determining the natural enemy complex or in estimating predation rates is unknown.

 

 

                                   3.  Genetic markers have been used to track parasitoids and assess their impact against hosts, such as common muscoid flies.  Legner & Brydon (1966) liberated a thelytokous race of parasitoid on poultry farms which they were able to tract and derive host mortality data from.   Legner et al. (1990, 1991) derived similar information by releasing gregarious strains of Muscidifurax raptorellus Kogan & Legner, and a temporary interference of several weeks with resident parasitism during the establishment phase was detected.  However, this was later overcome when the released strain had a chance to multiply naturally at the site.

 

                 D.  Visual Counts.

 

  1.  There are several advantages of using visual counts over many of the exclusion techniques.  There is no manipulation of the environment required.  Prey can be added or predators removed to determine the response of the predator to changes in prey density.  A visual record reveals the predator's diet in the field.  Perhaps serology and electrophoresis share these three advantages, but the latter require considerable technology.  visual counts require a substantial commitment of time to observe the predation and to determine the feeding rates for different combinations of predator/prey stages. 

 

  2.  Vision cannot be used if the predator is cryptic, easily disturbed or escapes from the observer.  Also, the time a predators spends consuming a prey may vary depending on the range of prey stages attacked, the hunger level of the predator, interference or stimulation by other predators or prey that are active in the area, and differences among individual predators due to genera, reproductive stage or molting (Luck et al. 1999).  These in turn can determine the probability that a predator will be observed in the field with a prey.  Laboratory data on the time spent by four predators consuming prey was highly variable, leaving the investigators pessimistic about the visual method's utility for estimating predation rates (Kiritani et al. 1972).  But the approach may still be valid for some predators, and has been used to determine the fraction of diurnal predation for each predator species in a complex (Elvin et al. 1983).

 

 

E.  Statistical Sampling.

 

                                    1.  Obtaining field samples during parasitoid/predator liberation periods can provide useful information about the ability of a species to effect its host/prey density.  Parasitoid impact was thus measured on the pink bollworm, Pectinophora gossypiella (Saunders).  A significant positive relationship was found between the total number of parasitoids released and the host density, which was most pronounced during a mid autumn period (Legner & Medved 1979).  Releases of egg-larval and larval-parasitoids produced small measurable reductions in P. gossypiella moth emergence from mature cotton bolls, but did not significantly reduce % boll infestation.  By releasing parasitoids at three densities, it was possible to show significant differences between controls and a low and medium release rate, but excessive parasitoid dispersal out of the release areas into the cotton field at large explained a leveled slope after the medium release rate (Legner & Medved 1979). 

 

  2.  The potential of Goniozus spp. and Pentalitomastix plethorica Caltagirone to regulate navel orangeworm, Amyelois transitella (Walker) was judged from seasonal positive functional responses to host density and with k-value analyses (Varley et al. 1974, Legner & Silveira-Guido 1983).  Application of this technique requires that the host show minimal overlapping of generations, however.  Goniozus emigratus (Rohwer) and Goniozus legneri Gordh demonstrated a significant capacity to recognize and respond in a regulative fashion in mid summer by increasing attack rates on higher host densities.  However, no such tendency was indicated during cooler periods of late autumn. 

 

                                  3.  Indigenous parasitism of Rhagoletis completa Cresson in its native range of western Texas and southeastern New Mexico was also assessed with k-value analysis,which showed a significant impact of combined natural mortality on host reduction (Legner & Goeden 1987).  Biosteres sublaevis Wharton demonstrated the greatest measurable activity as a cause of natural mortality.

 

  4.  Legner & Brydon (1966) were able to show an increased parasitism and house fly host mortality closer to liberation sites of parasitoids.  Legner et al. (1990, 1991) also charted increases and spread of muscoid fly parasitism from release sites.  The importance of proper field sampling, measurement of host destruction and unpredicted upsets to organisms in different guilds in these and similar studies was emphasized (Legner & Bay 1964, Legner 1979, Legner 1983, 1986).

 

                  5.  Long term sampling of the width of aquatic weed masses following the introduction of phytophagous cichlid fish established a regulatory capacity for Tilapia zillii Gervais in California irrigation canals (Legner & Fisher 1980, Legner & Murray 1981).  The data clearly showed seasons of high, medium and low regulatory capacity.

 

REFERENCES:

 

Ashby, J. W.  1974.  A study of arthropod predation of Pieris rapae L. using serological and exclusion techniques.  J. Appl. Ecol. 11:  419-25.

 

Aveling, C.  1981.  The role of Anthocoris species (Hemiptera: Anthocoridae) in the integrated control of the Damson-hop aphid (Phorodon humuli).  Ann. Appl. Biol. 97:  143-53.

 

Barry, R. M., J. H. Hatchett & R. D. Jackson.  1984.  Cage studies with predators of the cabbage looper, Trichoplusia ni, and corn ear worm, Heliothis zea in soybeans.  J. Georgia Ent. Soc. 9:  71-8.

 

Bartlett, B. R.  1968.  Outbreaks of two-spotted spider mites and cotton aphids following pesticide treatment.  I.  Pest stimulation vs. natural enemy destruction as the cause of outbreaks.  J. Econ. Ent. 61:  297-303.

 

Baumgaertner, J. U., A. P. Gutierrez & C. G. Summers.  1981.  The influence of aphid prey consumption on searching behavior, weight increase, developmental time and mortality of Chrysopa carnea (Neuroptera: Chrysopidae) and Hippodamia convergens (Coleoptera: Coccinellidae) larvae.  Canad. Ent. 113:  1007-14.

 

Bellows, T. S., Jr., R. G. Van Driesche & J. S. Elkinton.  1989b.  Extensions to Southwood and Jepson's graphical method of estimating numbers entering a stage for calculating losses to parasitism.  Res. Popul. Ecol.  Res. Popul. Ecol. 31:  169-84.

 

Boreham, P. F. L. & C. E. Ohiagu.  1978.  The use of serology in evaluating invertebrate predator-prey relationships:  a review.  Bull. Ent. Res. 68:  171-94.

 

Brown, A. W. A.  1951.  Insect Control by Chemicals.  John Wiley & Sons, New York.  817 p.

 

Bull, C. G. & M. V. King.  1923.  The identification of the blood meal of mosquitoes by means of the precipitin test.  Amer. S. Hyg. 3:  491-96.

 

Buonaccorsi, J. P. & J. S. Elkinton.  1990.  Estimation of contemporaneous mortality factors.  Res. Popul. Ecol. 32:  1-21.

 

Campbell, C. A. M.  1978.  Regulation of the Damson-hop aphid (Phorodon humuli [Schrank]) on hops (Humulus lupulus Linnaeus) by predators.  J. Hort. Sci. 53:  235-42.

 

Carl, K. P.  1982.  Biological control of native pests by introduced natural enemies.  Biol. Cont. News Info. 3:  190-200.

 

Carter,  N., I. F. G. McLean, A. E. Watt & A. G. G. Dixon.  1980.  Cereal aphids:  a case study and review. p. 272-349.  In:  T. H. Cooker (ed.), Applied Biology.  Academic Press, London.

 

Castanera, P., H. D. Loxdale & K. Novak.  1983.  Electrophoretic study of enzymes from cereal aphid populations. II.  Use of electrophoresis for identifying aphidiid parasitoids (Hymenoptera) of Sitobion avenae (Fabricius) (Hemiptera: Aphididae).  Bull. Ent. Res. 73:  659-65.

 

Chambers, R. J., K. D. Sunderland, I. J. Wyatt & G. P. Vickerman.  1983.  The effects of predator exclusion and caging on cereal aphids in winter wheat.  J. Appl. Ecol. 20:  209-24.

 

Charboussou, F.  1965.  La multiplication par vole trophique des tetrayques a la suite des traitements pesticides.  Relations avec les phenomenes de resistance acquise.  Boll. Zool. Agrar. Bachic. 7:  143-84.

 

Chelliah, S., L. Fabellar & A. E. Heinrichs.  1980.  Effects of sublethal doses of three insecticides on the reproductive rate of the brown planthopper, Nalaparvata lugens, on rice.  Environ. Ent. 9:  778-80.

 

CIBC.  1971.  Biological control programs against insects and weeds in Canada, 1959-1968.  Commonw. Agric. Bur. Tech. Comm., Vol. 4.  266 p.

 

Clausen, C. P. (ed.).  1978.  Introduced Parasites and Predators of Arthropod Pests and Weeds:  a World Review.  U. S. Dept. Agric. Handbk. 480.  545 p.

 

Coaker, T. H.  1965.  Further experiments on the effects of beetle predation on the numbers of cabbage root fly, Eriocschia brassicae (Bouché) attacking brassica crops.  Ann. Appl. Biol. 56:  7-20.

 

Cochran, W. G.  1963.  Sampling Techniques, 2nd Ed.  John Wiley & Sons, New York.

 

Cock, M. J. W.  1985.  A review of biological control of pests in the Commonwealth Caribbean and Bermuda up to 1982.  Commonw. Agric. Bur. Tech. Commun., Vol. 9.  218 p.

 

Croft, B. A. & A. W. A. Brown.  1975.  Responses of arthropod natural enemies to insecticides.  Ann. Rev. Ent. 20:  285-335.

 

Crook, N. E. & K. D. Sunderland.  1984.  Detection of aphid remains in predators by ELISA.  Ann. Appl. Biol. 105:  413-22.

 

DeBach, P.  1946.  An insecticidal check method for measuring the efficacy of entomophagous insects.  J. Econ. Ent. 39:  695-97.

 

DeBach. P.  1955.  Validity of insecticidal check method as a measure of the effectiveness of natural enemies of diaspine scale insects.  J. Econ. Ent. 48:  584-88.

 

DeBach, P. (ed.).  1964.  Biological Control of Insect Pests and Weeds.  Reinhold Publ. Corp., New York.

 

DeBach, P.  1974.  Biological Control by Natural Enemies.  Cambridge Univ. Press, London. 

 

DeBach, P. & C. B. Huffaker.  1971.  Experimental techniques for evaluation of the effectiveness of natural enemies. p. 113-40.  In:  C. B. Huffaker (ed.), Biological Control.  Plenum Press, New York.

 

DeBach, P., E. J. Dietrick & C. A. Fleschner.  1949.  A new technique for evaluating the efficiency of entomophagous insects in the field.  J. Econ. Ent. 42:  546.

 

Dempster, J. P.  1960.  A quantitative study of the predators on the eggs and larvae of the broom beetle, Phytodecta olivacea Forster, using the precipitin test.  J. Anim. Ecol. 29:  149-67.

 

Dempster, J. P.  1964.  The control of Pieris rapae with DDT 1.  The natural mortality of the young stages of Pieris.  J. Appl. Ecol. 4:  485-500.

 

Dempster, J. P.  1967.  The feeding habits of the Miridae (Heteroptera) living on broom (Sarothomnus scoparius L.).  Ent. Expt. Appl. 7:  149-54.

 

Dempster, J. P., O. W. Richards & N. Waloff.  1959.  Carabids as predators on the pupal stage of the chrysomelid beetle Phytodicta olivacea (Forster).  Oikos 10:  65-70.

 

Denno, R. F. & M. S. McClure (eds.).  1983.  Variable Plants and Herbivores in Natural and Managed Systems.  Academic Press, New York.

 

Dicke, M. & M. DeJong.  1986.  Prey preference of predatory mites:  electrophoretic analysis of the diet of Typhlodromus pyri Scheuten and Amblyseius finlandicus (Oudemans) collected in Dutch orchards.  Bull. IOBC/WPRS 9:  62-7.

 

Dittrich, V. P., P. Streibert & P. Bathe.  1974.  An old case reopened:  mite stimulation by insecticide residues.  Environ. Ent. 3:  534-39.

 

Dixon, A. F. G.  1959.  An experimental study of the search behavior of the predatory coccinellid beetle, Adalia compunctata (Linnaeus).  J. Anim. Ecol. 28:  259-81.

 

Dowden, P. B.  1962.  Parasites and Predators of Forest Insects Liberated in the United States Through 1960.  U. S. Dept. Agric. Handbk. 226.  70 p.

 

Downe, S. E. R. & A. S. West.  1954.  Progress in the use of precipitin test in entomological studies.  Canad. Ent. 86:  181-84.

 

Easteal, S. & I. A. Boussay.  1987.  A sensitive and efficient isozyme technique for small arthropods and other invertebrates.  Bull. Ent. Res. 77:  407-15.

 

Edmunds, G. F. & D. N. Alstad.  1985.  Malathion-induced sex ratio changes in black pine leaf scale (Hemiptera: Diaspididae).  Ann. Ent. Soc. Amer. 78:  403-05.

 

Ehler, L. E., K. G. Eveleens & R. van den Bosch.  1973.  An evaluation of some natural enemies of cabbage looper on cotton in California.  Environ. Ent. 2:  1009-15.

 

Ellington, J., K. Kiser, M. Cardenas, J. Duttle & Y. Lopez.  1984a.  The Insectavac:  A high-clearnace, high-volume arthropod vacuuming platform for agricultural ecosystems.  Environ. Ent. 13:  259-65.

 

Ellington, J., K. Kiser, G. Ferguson & M. Cardenas.  1984b.  A comparison of sweep-net, absolute and Insectavac sampling methods in cotton ecosystems.  J. Econ. Ent. 77:  599-605.

 

Elliot, J. M.  1977.  Some methods for the statistical analysis of samples of benthic invertebrates.  Sci. Publ. No. 25, Freshwater Biological Assoc.  Ferry House, United Kingdom.

 

Elvin, H. K., J. L. Stimac & W. H. Whitcomb.  1983.  Estimating rates of arthropod predation on velvetbean caterpillar larvae in soybeans.  Florida. Ent. 66:  319-30.

 

Eveleens, K. G., R. van den Bosch & L. E. Ehler.  1973.  Secondary outbreak induction of beet armyworm by experimental insecticide applications in cotton in California.  Environ. Ent. 2:  497-503.

 

Faeth, S. H. & D. Simberloff.  1981.  Population regulation of a leafmining insect, Cameraria sp. nov. at increased field densities.  Ecology 62:  620-24.

 

Falcon, L. A., R. van den Bosch, C. A. Ferris, L. K. Strombert, L, K. Etzel, R. E. Stinner & T. F. Leigh.  1968.  A comparison of season long cotton-pest-control programs in California during 1966.  J. Econ. Ent. 61:  633-42.

 

Ferro, D. N. & E. E. Southwick.  1984.  Microclimates of small arthropods:  estimating humidity within leaf boundary layer.  Environ. Ent. 13:  926-29.

 

Fichter, F. L. & W. P. Stephen.  1979.  Selection and use of host-specific antigens.  Misc. Publ. Ent. Soc. Amer. 11:  25-33.

 

Fichter, F. L. & W. P. Stephen.  1981.  Time related decay in prey antigens ingested by the predator Podisus maculiventris (Hemiptera: Pentatomidae) as detected by ELISA.  Oecologia 51:  404-07.

 

Fichter, F. L. & W. P. Stephen.  1984.  Time related decay in prey antigens ingested by arboreal spiders as detected by ELISA.  Environ. Ent. 13:  1583-87.

 

Flanders, S. E.  1942.  Absorptive development in parasitic Hymenoptera induced by food plant of the insect host.  J. Econ. Ent. 35:  834-35.

 

Fleschner, C. A.  1950.  Studies on searching capacity of the larvae of three predators of the citrus red mite.  Hilgardia 20:  233-65.

 

Fleschner, C. A.  1958.  Field approach to population studies of tetranychid mites in citrus and avocado in California.  Proc. 10th Internat. Congr. Ent. 2:  669-74.

 

Fleschner, C. A., J. C. Hall & D. W. Ricker.  1955.  Natural balance of mite populations in an avocado grove.  Calif. Avocado Soc. Yearbk. 39:  155-62.

 

Flint, M. L. & R. van den Bosch.  1981.  Introduction to integrated pest management.  Plenum Press, New York.

 

Folsom, J. W. & F. F. Bondy.  1930.  Calcium arsenate dusting as a cause of aphid infestations.  U. S. Dept. Agric. Cir. 116, Washington, D. C.  11 p.

 

Frank, J. H.  1967.  The insect predators of the pupal stage of the winter moth Operophtera brumata (Linnaeus) (Lepidoptera: Hydriomenidae).  J. Anim. Ecol. 36:  375-89.

 

Frazer, B. D. & N. Gilbert.  1976.  Coccinellids and aphids:  a quantitative study of the impact of adult ladybirds (Coleoptera: Coccinellidae) preying on field populations of pea aphids (Homoptera: Aphididae).  J. Ent. Soc. Brit. Columbia 73:  33-56.

 

Frazer, B. D. & B. Gill.  1981.  Hunger, movement and predation of Coccinella californica on pea aphids in the laboratory and in the field.  Canad. Ent. 113:  1025-33.

 

Frazer, B. D., N. Gilbert, N. Ives & D. A. Raworth.  1981a.  Predator reproduction and the overall predator-prey relationship.  Canad. Ent. 113:  1015-24.

 

Frazer, B. D., N. Gilbert, V. Nealis & D. A. Raworth.  1981b.  Control of aphid density by a complex of predators.  Canad. Ent. 113:  1035-41.

 

Futuyama, D. J. & S. C. Peterson.  1985.  Genetic variation in the use of resources by insects.  Ann. Rev. Ent. 30:  217-38.

 

Gardner, W. A., M. Shepard & R. Noblet.  1981.  Precipitin test for examining predator-prey interactions in soybean fields.  Canad. Ent. 113:  365-69.

 

Gates, D. M.  1980.  Biophysical Ecology.  Springer-Verlag, New York.

 

Gilbert, N., A. P. Gutierrez, B. Frazer & R. E. Jones.  1976.  Ecological Relationships.  Freeman & Sons, San Francisco.  156 p.

 

Giller, P. S.  1982.  The natural diets of waterbugs (Hemiptera: Heteroptera):  electrophoresis is a potential method of analysis.  Ecol. Ent. 7:  233-37.

 

Giller, P. S.  1984.  Predator gut state and prey detectability using electrophoretic analysis of gut content.  Ecol. Ent. 9:  157-62.

 

Giller, P. S.  1986.  The natural diet of the Notonectidae:  field trials using electrophoresis.  Ecol. Ent. 11:  163-72.

 

Gonzalez, D., D. A. Ramsay, T.a F. Leigh, B. S. Ekhom & R. van den Bosch.  1977.  A comparison of vacuum and whole-plant methods for sampling predaceious arthropods in cotton.  J. Econ. Ent. 66:  750-60.

 

Gould, J. R.  1990.  Estimating the impact of parasitoids on the dynamics of populations of gypsy moths.  Ph.D. dissertation, University of Massachusetts, Amherst, Mass.

 

Gould, J. R., R. G. Van Driesche, J. S. Elkinton, T. S. Bellows, Jr. & T. M. O'Dell.  1989.  A review of techniques for measuring the impact of parasitoids of lymantriids. p. 517-31.  In:  The Lymantriidae: Comparisons of Features of New and Old World Tussock Moths.  Northeast Forest Service Expt. Sta., General Tech. Rept. NE-123, Broomall, Pennsylvania.

 

Gould, J. R., J. S. Elkinton & W. E. Wallner.  1990a.  Density-dependent suppression of experimentally created gypsy moth, Lymantria dispar (Lepidoptera: Lymantriidae), populations by natural enemies.  J. Anim. Ecol. 59:  213-33.

 

Greathead, D. S.  1971.  A Review of Biological Control in the Ethiopian Region.  Commonw. Agric. Bur. Tech. Commun., Vol. 5.  1962 p.

 

Greathead, D. S. (ed.).  1976.  A Review of Biological Control in Western and Southern Europe.  Commonw. Agric. Bur. Tech. Commun., Vol. 7.  182 p.

 

Green, R. H.  1979.  Sampling Design and Statistical Methods For Environmental Biologists.  John Wiley & Sons, New York. 

Greenstone, M. H.  1977.  A passive haemagglutination inhibition assay for the identification of stomach contents of invertebrate predators.  J. Appl. Ecol. 14:  457-64.

 

Greenstone, M. H.  1979.  Passive haemagglutin inhibition:  a powerful new tool for field studies of entomophagous predators.  Misc. Publ. Ent. Soc. Amer. 11:  69-78.

 

Greenstone, M. H.  1983.  Amblyospora site-specificity and site-tenacity in a wolf spider:  a serological dietary analysis.  Oecologia 56:  79-83.

 

Guerin, P. M. & E. Städler.  1984.  Carrot fly cultivar preferences:  some influencing factors.  Ecol. Ent. 9:  413-20.

 

Hafez, M.  1961.  Seasonal fluctuations of population density of the cabbage aphid, Brevicoryne brassicae (L.), in the Netherlands, and the role of its parasite, Aphidius (Diaeretiella) rapae (Curtis).  Neth. J. Plant Pathol. 67:  445-548.

 

Hall, R. R., A. E. R. Downe, C. R. McClellan & A. S. Wiest.  1953.  Evaluation of insect predator-prey relationships by precipitin stest studies.  Mosq. News 13:  199-204.

 

Hand, L. F. & A. J. Keaster.  1967.  The environment of an insect field cage.  J. Econ. Ent. 60:  910-15.

 

Hare, J. D. 1990.  Effects of plant variation on herbivore-natural enemy interactions.  In:  R. S. Fritz & E. L. Simms (eds.), Ecology and Evolution of Plant Resistance.  Univ. of Chicago Press, Chicago.

 

Hassell, M. P.  1980.  Foraging strategies, population models and biological control:  a case study.  J. Anim. Ecol. 49:  603-28.

 

Hawks, R. B.  1972.  A fluorescent dye technique for marking insect eggs in predation studies.  J. Econ. Ent. 65:  1477-1478.

 

Hazzard, R., D. N. Ferro & R. G. Van Driesche.  1991.  Impact of Coleomegilla maculata (Coleoptera: Coccinellidae) and other native predators on eggs of the Colorado potato beetle, Leptinotarsa decemlineata (Say) (Coleoptera: Chrysomelidae) in potatoes in Massachusetts.  Environ. Ent. 20.

 

Höller, C. & J. J. Braune.  1988.  The use of isoelectric focusing to assess percentage hymenopterous parasitism in aphid populations.  Ent. Exp. Appl. 47:  105-14.

 

Huffaker, C. B. & C. E. Kennett.  1966.  Studies of two parasites of the olive scale, Parlatoria oleae (Colvee).  IV.  Biological control of Parlatoria oleae (Colvée) through the compensatory action of two introduced parasites.  Hilgardia 37:  283-35.

 

Huffaker, C. B., C. E. Kennett & G. L. Finney.  1962.  Biological control of olive scale, Parlatoria oleae (Colvée) in California by imported Aphytis maculicornis (Masi) (Hymenoptera: Aphelinidae).  Hilgardia 32:  541-636.

 

Irwin, M. E., R. W. Gill & D. Gonzalez.  1974.  Field cage studies of native egg predators of the pink bollworm in southern California cotton.  J. Econ. Ent. 67:  193-96.

 

Ito, Y., H. Yamanaka, F. Nakasuji & K. Kiritani.  1972.  Determination of predator-prey relationship with an activiable tracer, Europium-151.  Konchy 40:  278-83.

 

Ives, P. M.  1978.  How discriminating are cabbage butterflies?  Aust. J. Ent. 3:  261-76.

 

Jenkins, D. W.  1963.  Use of radionuclides in ecological studies of insects. p. 431-40.  In:  V. Schult & A. W. Klement (eds.), Radioecology.  Rheinhold Publ. Co., New York.

 

Jenkins, D. W. & C. C. Hassett.  1950.  Radioisotopes in entomology.  Nucleonics 6:  5-14.

 

Jessen, R. L.  1978.  Statistical Survey Techniques.  John Wiley & Sons, New York. 

 

Jones, V. P. & M. P. Parrella.  1984.  The sublethal effects of selected pesticides on life table parameters of Panonychus citri (Acari: Tetranychidae).  Canad. Ent. 116:  1033-40.

 

Jones, V. P., R. R. Youngman & M. P. Parrella.  1983.  The sublethal effects of selected insecticides on photosynthetic rates of lemon and orange leaves in California.  J. Econ. Ent. 76:  1178-80.

 

Karieva, P.  1985.  Patchiness, dispersal, and species interactions:  consequences for communities of herbivorous insects. p. 192-206.  In:  T. Case & J. Diamond (eds.), Community Ecology.  Harper & Row, New York.

 

Keating, S. T., J. P. Burand & J. S. Elkinton.  1989.  DNA hybridization assay for detection of gypsy moth nuclear polyhedrosis virus in infected gypsy moth (Lymantria dispar L.) larvae.  Appl. Environ. Microbiol. 55:  2749-54.

 

Kelleher, J. S. & M. A. Hulme.  1984.  Biological control programmes against insects and weeds in Canada, 1969-1980.  Commonw. Agric. Bur. Tech. Commun., Vol. 8.  410 p.

 

Kenmore, P. E., F. D. Carino, C. A. Perez, V. A. Dyck & A. P. Gutierrez.  1984.  Population regulation of the rice brown planthopper (Nilaparvata lugens Stal) within rice fields in the Philippines.  J. Plant Prot. Trop. 1:  19-37.

 

Kinzer, R. E., C. B. Cowan, R. L. Ridgway, F. J. Davis, Jr. & R. J. Copperage.  1977.  Populations of arthropod predators and Heliothis spp. after applications of aldicarb and moncrotophos.  Environ. Ent. 6:  13-16.

 

Kirtitani, K. & F. Nakasuji.  1967.  Estimation of the stage-specific survival rate in the insect population with overlapping stages.  Res. Popul. Ecol. 9:  143-52.

 

Kiritani, K., S. Kawahera, T. Sasaba & F. Nakasuji.  1972.  Quantitative evaluation of predation by spiders on the green rice leafhopper, Nephotettix cincticeps Uhler, by a sight count method.  Res. Pop. Ecol. 13:  187-200.

 

Kolodny-Hirsch, D. M., R. C. Reardon, K. W. Thorpe & M. J. Raupp.  1988.  Evaluating the impact of sequential releases of Cotesia melanoscela (Hymenoptera: Braconidae) on Lymantria dispar (Lepidoptera: Lymantriidae).  Environ. Ent. 17:  403-08.

 

Kring, T. J., F. E. Gilstrap & F. J. Michels, Jr.  1985.  Role of indigenous coccinellids in regulating greenbugs (Homoptera: Aphididae) on Texas grain sorghum.  J. Econ. Ent. 78:  269-73.

 

Laing, J. E. & J. Hamai.  1976.  Biological control of insect pests and weeds by imported parasites and predators. p. 685-743.  In:  C. B. Huffaker & P. S. Messenger (eds.), Theory and Practice of Biological Control.  Academic Press, New York. 788 p.

 

Legner, E. F.  1967.  Behavior changes the reproduction of Spalangia cameroni, S. endius, Muscidifurax raptor, and Nasonia vitripennis (Hymenoptera: Pteromalidae) at increasing fly host densities.  Ann. Ent. Soc. Amer. 60:  819-26.

 

Legner, E. F.  1969.  Distribution pattern of hosts and parasitization by Spalangia drosophilae (Hymenoptera; Pteromalidae).  Canad. Ent. 101:  551-7.

 

Legner, E. F.  1971.  Some effects of the ambient arthropod complex on the density and potential parasitization of muscoid Diptera in poultry wastes.  J. Econ. Ent. 64:  111-15.

 

Legner, E. F.  1979.  The relationship between host destruction and parasite reproductive potential in Muscidifurax raptor, M. zaraptor, and Spalangia endius [Chalcidoidea: Pteromalidae].  Entomophaga 24(2):  145-152.

 

Legner, E. F.  1983.  Requirements for appraisal of the role of parasitic insects in the natural control of synanthropic Diptera.  Proc. Calif. Mosq. & Vect. Contr. Assoc., Inc. 51:  97-8.

 

Legner, E. F.  1986.  The requirement for reassessment of interactions among dung beetles, symbovine flies and natural enemies.  Ent. Soc. Amer. Misc. Publ. 61:  120-31.

 

Legner, E. F. & E. C. Bay.  1964.  Natural exposure of Hippelates eye gnats to field parasitization and the discovery of one pupal and two larval parasites.  Ann. Entomol. Soc. Amer. 57(6):  767-769.

 

Legner, E. F. & H. W. Brydon.  1966. Suppression of dung inhabiting fly populations by pupal parasites.  Ann. Entomol. Soc. Amer. 59(4):  638-651.

 

Legner, E. F. & T. W. Fisher.  1980.  Impact of Tilapia zillii (Gervais) on Potamogeton pectinatus L., Myriophyllum spicatum var. exalbescens Jepson, and mosquito reproduction in lower Colorado Desert irrigation canals.  Acta. Oecologica, Oecol. Applic. 1:  3-14.

 

Legner, E. F. & D. Gerling.  1967.  Host-feeding and oviposition on Musca domestica by Spalangia cameroni, Nasonia vitripennis, and Muscidifurax raptor (Hymenoptera: Pteromalidae) influences their longevity and fecundity.  Ann. Ent. Soc. Amer. 60:  678-91.

 

Legner, E. F. & R. D. Goeden.  1987.  Larval parasitism of Rhagoletis completa (Diptera: Tephritidae) on Juglans microcarpa (Juglandaceae) in western Texas and southeastern New Mexico.  Proc. Ent. Soc. Wash. 89:  739-43.

 

Legner, E. F. & R. A. Medved.  1979.  Influence of parasitic Hymenoptera on the regulation of pink bollworm, Pectinophora gossypiella, on cotton in the Lower Colorado Desert.  Environ. Ent. 8:  922-30.

 

Legner, E. F. & C. A. Murray.  1981.  Feeding rates and growth of the fish Tilapia zillii [Cichlidae] on Hydrilla verticillata, Potamogeton pectinatus and Myriophyllum spicatum var. exalbescens and interactions in irrigation canals of southeastern California.  J. Amer. Mosq. Contr. Assoc. 41:  241-50.

 

Legner, E. F. & A. Silveira-Guido.  1983.  Establishment of Goniozus emigratus and Goniozus legneri [Hym: Bethylidae] on navel orangeworm, Amyelois transitella [Lep: Phycitidae] in California and biological control potential.  Entomophaga 28:  97-106.

 

Legner, E. F., W. D. McKeen & R. W. Warkentin.  1990.  Inoculation and spread of parasitic wasps to control filth flies in poultry houses.  Proc. Calif. Mosq. & Vect. Contr. Assoc., Inc. 57:  146-50.

 

Legner, E. F., W. D. McKeen & R. W. Warkentin.  1991.  Inoculation of three pteromalid wasp species (Hymenoptera: Pteromalidae) increases parasitism and mortality of Musca domestica L. pupae in poultry manure.  Bull. Soc. Vect. Ecol. 16:

 

Lingren, P. D., R. L. Ridgway & S. L. Jones.  1968.  Consumption by several common arthropod predators of eggs and larvae of two Heliothis species that attack cotton.  Ann. Ent. Soc. Amer. 61:  613-18.

 

Linit, M. J. & F. M. Stephen.  1983.  Parasite and predator component of within-tree southern pine beetle (Coleoptera: Scolytidae) mortality.  Canad. Ent. 115:  679-88.

 

Lister, A., M. B. Usher & W. Block.  1987.  Description and quantification of field attack rates by predator mites:  an example using an electrophoresis method with a species of anarchic mite.  Oecologia 72:  185-91.

 

Luck, R. F.  1981.  Parasitic insects introduced as biological control agents for arthropod pests. p. 125-284.  In:  D. Pimentel (ed.), CRC Handbook of Pest Management in Agriculture.  CRC Press, Boca Raton, Florida.  504 p.

 

Luck, R. F. & D. L. Dahlsten.  1975.  Natural decline of a pine needle scale (Chionaspis pinifoliae [Fitch]) outbreak at South Lake Tahoe, California, following cessation of adult mosquito control with malathion.  Ecology 56:  983-904.

 

Luck, R. F., R. van den Bosch & R. Garcia.  1977.  Chemical insect control, a troubled pest management strategy.  BioScience 27:  606-11.

 

Luck, R. F., J. C. van Lenteren, P. H. Twine, L. Kuenen & T. Unruh.  1979.  Prey or host searching behavior that leads to a sigmoid functional response in invertebrate predators and parasitoids.  Res. Popul. Ecol. 20:  257-64.

 

Luck, R. F., B. M. Shepard & P. E. Kenmore.  1999..  Evaluation of biological control with experimental methods.  In:  Principles and Application of Biological Control.  Academic Press, San Diego, CA.  1046 p.

 

MacPhee, A. W. & C. R. MacLellan.  1971.  Ecology of apple orchard fauna and development of integrated pest control in Nova Scotia.  Proc. Tall Timbers Conf. on Ecol. Animal Control Habitat Management 3:  197-208.  Tallahassee, Florida, Tall Timbers Res. Station.

 

Maggi, V. L. & T. F. Leigh.  1983.  Fecundity response of the two-spotted spider mite to cotton treated with methyl parathion or phosphoric acid.  J. Econ. Ent. 76:  20-25.

 

Manly, B. F. J.  1974.  Estimation of stage-specific survival rates and other parameters for insect populations developing through several life stages.  Oecologia 15:  277-85.

 

Manly, B. F. J.  1976.  Extensions to Kiritani and Nakasuji's method for analysing insect stage-frequency data.  Res. Popul. Ecol. 17:  191-99.

 

Manly, B. F. J.  1977.  A further note on Kiritani and Nakasuji's model for stage-frequency data including comments on the use of Tukey's jackknife technique for estimating variances.  Res. Popul. Ecol. 18:  177-86.

 

Manly, B. F. J.  1989.  Analysis of stage-frequency data. p. 3-69.  In:  L. McDowell (ed.), Esimtation and Analyses of Insect Populations.  Springer Verlag, New York.

 

Mattson, W. J., Jr.  1980.  Herbivory in relation to plant nitrogen content.  Ann. Rev. Ecol. Syst. 11:  119-61.

 

Matsumoto, B. M. & T. Nishida.  1966.  Predator-prey investigations of the taro leafhopper and its egg predator.  Hawaii Agr. Expt. Sta. Tech. Bull. 64, Univ. of Hawaii, Honolulu.  32 p.

 

McCarty, M. T., M. Shepard & S. G. Turnipseed.  1980.  Identification of predaceous arthropods in soybeans by using autoradiography.  Environ. Ent. 9:  199-203.

 

McDaniel, S. G. & W. L. Sterling.  1979.  Predator determination and efficiency of Heliothis virescens eggs in cotton using 32P2. Environ. Ent. 8:  1083-87.

 

McGuggan, B. M. & H. C. Coppel.  1962.  Biological control of forest insects 1910-1058.  p. 35-127.  In:  A Review of the Biological Control Attempts Against Insects and Weeds in Canada.  Commonw. Agric. Bur. Tech. Commun., Vol. 2.

 

McGuire, M. R. & J. E. Henry.  1989.  Production and partial characterization of monoclonal antibodies for detection of entomopoxvirus from Melanoplus sanguinipes.  Ent. Expt. Appl. 51:  21-8.

 

McLeod, J. H.  1962.  Biological control of pests of crops, fruit trees, ornamentals and weeds in Canada up to 1959.  p. 1-33.  In:  A Review of the Biological Control Attempts Against Insects and Weeds in Canada.  Commonw. Agric. Bur. Tech. Commun., Vol. 2.

 

Metcalf, R. L. & W. Luckmann.  1982.  Introduction to Insect Pest Management.  John Wiley & Sons, New York.

 

Miller, M. C. (ed.).  1979.  Serology in Insect Predator-prey Studies.  Misc. Publ. Ent. Soc. Amer. 11(4):  1-84.

 

Morris, R. F.  1955.  The development of sampling techniques for forest defoliators with particular reference to the spruce budworm.  Canad. J. Zool. 33:  225-94.

 

Morris, R. F.  1960.  Sampling insect populations.  Ann. Rev. Ent. 5:  243-64.

 

Moulder, B. C. & D. E. Reichle.  1972.  Significance of spider predation in the energy dynamics of forest-floor arthropod communities.  Ecol. Monogr. 42:  473-98.

 

Mueke, J. M., G. R. Manglitz & W. R. Kerr.  1978.  Pea aphid:  interaction of insecticides and alfalfa varieties.  J. Econ. Ent. 71:  61-65.

 

Murray, R. A. & M. G. Solomon.  1978.  A rapid technique for analyzing diets of invertebrate predators by electrophoresis.  Ann. Appl. Biol. 90:  7-10.

 

Myers, J. H.  1985.  Effect of physiological condition of the host plant on the ovipositional choice of the cabbage white butterfly, Pieris rapae.  J. Anim. Ecol. 54:  193-204.

 

Nordlund, D. A., R. L. Jones & W. J. Lewis (eds.).  1981.  Semiochemicals:  Their Role in Pest Control.  John Wiley & Sons, New York. 

 

Oatman, E. R. & G. R. Platner.  1971.  Biological control of the tomato fruitworm, cabbage looper, and hornworms on processing tomatoes in southern California, using mass releases of Trichogramma pretiosum.  J. Econ. Ent. 64:  501-06.

 

Oatman, E. R. & G. R. Platner.  1978.  Effects of mass releases of Trichogramma pretiosum against lepidopterous pests on processing tomatoes in southern California, with notes on host egg population trends.  J. Econ. Ent. 71:  896-900.

 

Ohiagu, C. E. & P. F. L. Boreham.  1978.  A simple field test for evaluating insect prey-predator relationships.  Ent. Expt. Appl. 23:  40-47.

 

Olton, G. S. & E. F. Legner.  1975.  Winter inoculative releases of parasitoids to reduce houseflies in poultry manure.  J. Econ. Entomol. 68(1):  35-38.

 

Papaj, D. R. & M. D. Rausher.  1987.  Components of conspecific host discrimination behavior in the butterfly, Battus philenor.  Ecology 68:  245-53.

 

Pendleton, R. C. & A. W. Grundmann.  1954.  Use of 32P in tracing some insect plant relationships of the thistle, Cirsium undulatum.  Ecology 35:  187-91.

 

Pickett, A. D. & N. A. Patterson.  1953.  The influence of spray programs on the fauna of apple orchards in Nova Scotia.  IV.  A review.  Canad. Ent. 85:  472-78.

 

Price, P. W., C. E. Bouton, P. Gross, B. A. McPheron, J. N. Thompson & A. E. Weis.  1980.  Interactions among three trophic levels:  influence of plants on interactions between insect herbivores and natural enemies.  Ann. Rev. Ecol. Syst. 11:  41-65.

 

Rabbinge, R., G. W. Ankersmit & G. A. Pak.  1979.  Epidemiology and simulation of population development of Sitobion avenae in winter wheat.  Neth. J. Plant Pathol. 85:  197-220.

 

Ragsdale, D. W., A. D. Larson & L. D. Newson.  1981.  Quantitative assessment of the predators of Nezara viridula eggs and nymphs within a soybean agroecosystem using an ELISA.  Environ. Ent. 10:  402-05.

 

Rao, V. P., M. A. Ghani, T. Sankaran & K. C. Mather.  1971.  A review of the biological control of insects and other pests in Southeast Asia and the Pacific region.  Commonw. Agric. Bur. Tech. Commun., Vol. 6.  149 p.

 

Readshaw, J. L.  1973.  The numerical response of predators to prey density.  J. Appl. Ecol. 10:  342-51.

 

Ressig, W. H., E. A. Heinfichs & S. L. Valencia.  1982.  Insecticide induced resurgence of the brown plant hopper, Nilaparvata lugens, on rice varieties with different levels of resistance.  Environ. Ent. 11:  165-68.

 

Richards, O. W. & N. Waloff.  1961.  A study of a natural population of Phytodecta olivacea (Forester) (Coleoptera, Chrysomelidae).  Philos. Trans. Roy. Soc. London Ser. B 244:  205-57.

 

Richards, O. W., N. Waloff & J. P. Spradberry.  1960.  The measurement of mortality in an insect population in which recruitment and mortality widely overlap.  Oikos 11:  306-10.

 

Richman, D. B., R. C. Hemenway, Jr. & W. H. Whitcomb.  1980.  Field-cage evaluation of predators of the soybean looper, Pseudoplusia includens (Lepidoptera: Noctuidae).  Environ. Ent. 9:  315-17.

 

Ridgway, R. L. & S. B. Vinson.  1976.  Biological control by augmentation of natural enemies:  insect and mite control with parasites and predators.  Plenum Press, New York.

 

Ripper, W. E.  1956.  Effect of pesticides on balance of arthropod populations.  Ann. Rev. Ent. 1:  403-38.

 

Rombach, M. C., R. M. Aguda, B. M. Shepard & D. W. Roberts.  1986a.  Entomopathogenic fungi (Deuteromycotina) in the control of the black bug of rice, Scotinophara coarctata (Hemiptera: Pentatomidae).  J. Invert. Path. 48:  174-79.

 

Rombach, M. C., R. M. Aguda, B. M. Shepard & D. W. Roberts.  1986b.  Infection of rice brown planthoppers, Nilaparvata lugens (Homoptera: Delphacidae) by field application of entomopathogenic hyphomycetes (Deuteromycotina).  Environ. Ent. 15:  1070-73.

 

Room, P. M.  1987.  32P labelling of immature stages of Heliothis armigera (Hübner) and H. punctigera Wallengren (Lepidoptera: Noctuidae): relationships of doses to radioactivity, mortality and label half-life.  J. Aust. Ent. Soc. 16:  245-51.

 

Rosen, D. & P. DeBach.  1978.  Diaspididae. p. 78-128.  In:  C. P. Clausen (ed.), Introduced Parasites and Predators of Arthropod Pests and Weeds:  A World Review.  U. S. Dept. Agric. Handbk 480.

 

Rothchild, G. H. L.  1966.  A study of natural populations of Conomelus anceps (Germer) (Homoptera: Delphacidae) including observations on predation using the precipitin test.  J. Anim. Ecol. 35:  413-34.

 

Rothchild, G. H. L.  1970.  Observations on the ecology of the rice ear bug, Leptocoris oratorius (Hemiptera: Alydidae) in Sarawak (Malaysian Borneo).  J. Appl. Ecol. 7:  147-67.

 

Rothchild, G. H. L.  1971.  The biology and ecology of the rice stem borer in Sarawak (Malaysian Borneo).  J. Appl. Ecol. 8:  287-332.

 

Ruesink, W. G.  1975.  Estimating time-varying survival of arthropod life stages from population density.  Ecology 56:  244-47.

 

Russell, D. A.  1987.  A simple method for improving estimates of percentage parasitism by insect parasitoids from field sampling of hosts.  New Zealand Ent. 10:  38-40.

 

Ryan, R. B. & R. D. Medley.  1970.  Test release of Itoplectis quadricingulatus against European pine shoot moth in an isolated infestation.  J. Econ. Ent. 63:  1390-92.

 

Sabelis, M. W.  1981.  Biological control of two-spotted spider mites using phytoseiid predators.  Part I.  Modelling the predator-prey interactions at the individual level.  Ph.D. Thesis, Agric. Univ. Wageningen, The Netherlands.

 

Sailer, R. I.  1966.  An aphid predator exclusion test.  p. 263.  In:  J. Hodek (ed.), Ecology of Aphidiophagous Insects.  Junk, The Hague, Netherlands.

 

Schneider, B., H. Podoler & D. Rosen.  1988.  Population dynamics of the Florida wax scale, Ceroplastes floridensis (Homoptera: Coccidae), on citrus in Israel. 5.  Effect of the parasite, Tetrastichus ceroplastae (Girault).  Acta Oecol. 9:  75-83.

 

Scriber, J. M. & F. Slansky, Jr.  1981.  The nutritional ecology of immature insects.  Ann. Rev. Ent. 26:  183-211.

 

Shaw, J. G., R. P. Sunderman, D. S. Moreno & D. S. Fargerlund.  1973.  California red scale:  females isolated by treatments with dichlorvos.  Environ. Ent. 2:  1062.

 

Shepard, B. M. & V. H. Waddill.  1976.  Rubidium as a marker for Mexican bean beetles, Epilachna varivestis (Coleoptera: Coccinellidae).  Canad. Ent. 108:  337-39.

 

Shepard, B. M. & G. S. Arida.  1986.  Parasitism and predation of yellow stemborer Scirpophaga incertulus (Walker) (Lepidoptera: Pyralidae) eggs in transplanted and direct seeded rice.  J. Ent. Sci. 21:  26-32.

 

Smith, H. S. & P. DeBach.  1942.  The measurement of the effect of entomophagous insects on population densities of their hosts.  J. Econ. Ent. 4:  231-34.

 

Smith, R. F. & R. van den Bosch.  1967.  Integrated pest management. p. 295-340.  In:  W. W. Kilgore & R. L. Doutt (eds.), Pest Control.  Academic Press, New york.  477 p.

 

Soop, P. I. & K. D. Sunderland.  1989.  Some factors affecting the detection of aphid remains in predators using ELISA.  Entomol. Exp. Appl. 51:  11-20.

 

Southwood, T. R. E.  1978.  Ecological Methods With Particular Reference to the Study of Insect Populations.  John Wiley & Sons, New York.

 

Sparks, A. N., H. C. Chaing, G. T. Burkhardt, M. L. Fairchild & G. T. Weekman.  1966.  Evaluation of the influence of predation on corn borer populations.  J. Econ. Ent. 59:  104-07.

 

Starks, K. J., R. Muniappan & R. D. Eikenbary.  1972.  Interaction between plant resistance and parasitism against the greenbug on barley and sorghum.  Ann. Ent. Soc. Amer. 65:  650-55.

 

Stern, V. M., R. F. Smith, R. van den Bosch & K. S. Hagen.  1959.  The integration of chemical and biological control of the spotted alfalfa aphid.  The integrated control concept.  Hilgardia 29:  81-101.

 

Stimmann, M. W.  1974.  Marking insects with rubidium:  imported cabbage worm marked in the field.  Environ. Ent. 3:  327-28.

 

Stuart, A.  1976.  Basic Ideas of Scientific Sampling.  Griffin's Statistical Monograph No. 4.  Hafner Press, New York.

 

Sunderland, K. D.  1988.  Quantitative methods for detecting invertebrate predation occurring in the field.  Ann. Appl. Biol. 112:  201-224.

 

Sunderland, K. D. & S. L. Sutton.  1980.  A serological study of arthropod predation on wood lice in a dune grassland ecosystem.  J. Anim. Ecol. 49:  987-1004.

 

Sunderland, K. D., N. W. Crook, D. L. Stacy & B. J. Fuller.  1987.  A study of feeding by polyphagous predators or cereal aphids using ELISA and gut dissections.  J. Appl. Ecol. 907_33.

 

Thompson, W. R.  1955.  Mortality factors acting in a sequence.  Canad. Ent. 87:  264-75.

 

Torgensen, T. R. & B. B. Ryan.  1981.  Field biology of Telenomus californicus Ashmead, and important egg parasitoid of Douglas fir tussock moth.  Ann. Ent. Soc. Amer. 74:  185-86.

 

Turnbull, A. L.  1964.  The search for prey by a web-building spider Achaearanea tepidariorum (C. L. Koch) (Araneae: Theridiidae).  Canad. Ent. 96:  568-79.

 

van den Bosch, R., T. F. Leigh, D. Gonzalez & R. E. Stinner.  1969.  Cage studies on predators of the bollworm in cotton.  J. Econ. Ent. 62:  1486-89.

 

van der Berg, H., B. M. Shepard, J. A. Litsinger & P. C. Pantua.  1988.  Impact of predators and parasitoids on the eggs of Rivulz atimeta, Naranga venescens (Lepidoptera: Noctuidae) and Hydrellia philippina (Diptera: Ephydridae) in rice.  J. Plant Prot. Trop. 5:  103-08.

 

van de Vrie, M. J., J. A. McMurtry & C. B. Huffaker.  1972.  Ecology of tetranychid mites and their natural enemies.  III.  Biology, ecology and pest status and host-plant relations of tetranychids.  Hilgardia 41:  343-432.

 

Van Driesche, R. G.  1983.  Meaning of "percent parasitism" in studies of insect parasitoids.  Environ. Ent. 12:  1611-22.

 

Van Driesche, R. G.  1988.  Measurement of population recruitment of Apanteles glomeratus (L.) (Hymenoptera: Braconidae), a parasitoid of Pieris rapae (L.) (Lepidoptera: Pieridae), and factors influencing adult parasitoid foraging success in kale. Bull. Ent. Res. 78: 199-208.

 

Van Driesche, R. G. & T. S. Bellows, Jr.  1988.  Host and parasitoid recruitment for quantifying losses from parasitism, with reference to Pieris rapae and Cotesia glomerata.  Ecol. Ent. 13:  215-22.

 

Van Driesche, R. G., T. S. Bellows, Jr., D. N. Ferro, R. Hazzard & M. Maher.  1989.  Estimating stage survival from recruitment and density data, with reference to egg mortality in the Colorado potato beetle, Leptinotarsa decemlineata (Say).  Canad. Ent. 121:  291-300.

 

Van Driesche, R. G., T. S. Bellows, Jr., J. S. Elkinton, J. Gould & D. N. Ferro.  1991a.  The meaning of percentage parasitism revisited:  solutions to the problem of accurately estimating total losses from parasitism.  Environ. Ent. 20:  1-7.

 

Van Driesche, R. G., D. N. Ferro, E. Carey & A. Maher.  1991b.  Assessing the impact of augmentative releases of parasitoids with the "recruitment method," with special reference to Edovum puttleri Grissel, and egg parasitoid of the Colorado potato beetle (Coleop.: Chrysomelidae).  Entomophaga.

 

van Lenteren, J. C. & K. Bakker.  1976.  Functional responses in invertebrates.  Neth. J. Zool. 26:  567-72. 

 

van Sickle, D. & R. M. Weseloh.  1974.  Habitat variables that influence the attack by hyperparasites of Apanteles melanoscelus cocoons.  J. New York. Ent. Soc. 82:  2-5.

 

Varley, G. C., G. R. Gradwell & M. P. Hassell.  1974.  Insect Population Ecology, an Analytical Approach.  Univ. of Calif. Press, Berkeley & Los Angeles.  212 p.

 

Vickermann, G. P. & K. D. Sunderland.  1975.  Arthropods in cereal crops:  nocturnal activity, vertical distribution and aphid predation.  J. Appl. Ecol. 12:  755-66.

 

Way, M. J.  1973.  Objectives, methods and scope of integrated control.  p. 137-52.  In:  P. W. Geier, L. R. Clark, D. J. Anderson & H. A. Nix (eds.).  Insects:  Studies in Population Management.  Mem. Ecol. Soc. Aust., Vol. 1.

Way, M. J. & C. J. Banks.  1958.  The control of Aphis fabae Scop. with special reference to biological control of insects which attack annual crops.  Proc. 10th Intern. Cong. Entomol. 4:  907-09.

 

Way, M. J. & C. J. Banks.  1968.  Population studies on the active stages of the black bean aphid Aphis fabae Scop., on its winter host Euonymus europaeus Linnaeus.  Ann. Appl. Biol. 62:  177-97.

 

Weseloh, R. M.  1974.  Host-related microhabitat preference of Gypsy moth larval parasite, Parasetigena agilis.  Environ. Ent. 3:  363-64.

 

Weseloh, R. M.  1978.  Seasonal and spatial mortality patterns of Apanteles melanoscelus due to predators and Gypsy moth hyperparasites.  Environ. Ent. 7:  662-65.

 

Weseloh, R. M.  1982.  Implications of tree microhabitat preferences of Compsilura concinata (Diptera: Tachinidae) for its effectiveness as a Gypsy moth parasitoid.  Canad. Ent. 114:  617-22.

 

Whitham, T. G., A. G. Williams & A. M. Robinson.  1984.  The variation principal, individual plants as temporal and spatial mosaics of resistance to rapidly evolving pests. p. 15-51.  In:  P. W. Price, C. N. Slobodchikoff & W. S. Gaud (eds.), A New Ecology--Novel Approaches to Interactive Systems.  John Wiley & Sons, New York.

 

Wilson, F.  1960.  A review of biological control of insects and weeds in Australia and Australian New Guinea.  Commonw. Agric. Bur. Tech. Commun., Vol. 1.  102 p.

 

Winder, L.  1990.  Predation of cereal aphid Sitobion avenae by polyphagous predators on the ground.  Ecol. Ent. 15:  105-10.

 

Woglum, R. S., J. R. LeFollette, W. E. Landon & H. C. Lewis.  1947.  The effect of field-applied insecticides on beneficial insects of citrus in California.  J. Econ. Ent. 40:  818-20.

 

Wolfson, J.  1980.  Oviposition responses of Pieris rapae to environmentally induced variation in Brassica nigra.  Ent. Expt. Appl. 27:  223-32.

 

Wood, B. J.  1971.  Development of integrated control program for tropical perennial crops in Malaysia. p. 113-40.  In:  C. B. Huffaker (ed.), Biological Control.  Plenum Press, New York.  511 p.

 

Wool, D., D. Gerling & I. Cohen.  1984.  Electrophoretic detection of two endoparasitic species, Encarsia lutea and Eretmocerus mundus in the whitefly, Bemisia tabaci (Genn.) (Hom., Aleurodidae).  Z. angew. Ent. 98:  276-79.

 

Wool, D., H. F. van Emden & S. W. Bunting.  1978.  Electrophoretic detection of the internal parasite, Aphidius matricariae in Myzus persicae.  Ann. Appl. Biol. 90:  21-6.

 

Wright, D. W., R. D. Hughes & J. Worell.  1960.  The effect of certain predators on the numbers of cabbage root fly (Erioschia brassicae [Bouché]) and on the subsequent damage caused by the pest.  Ann. Appl. Biol. 48:  756-63.