You are on page 1of 29

Agriculture, Ecosystems and Environment, 10 (1983) 247-274 247

Elsevier Science Publishers B. V., Amsterdam - Printed in The Netherlands

THE ECOLOGICAL BASIS OF BOLL WEEVIL (ANTHONOM US


GRANDIS BOHEMAN) MANAGEMENT

DALE G. BOTTRELL
University of California, Consortium for International Crop Protection, 2288 Fulton
Street, Suite 310, Berkeley CA 94704 (U.S.A.)
(Accepted 22 July 1983)

ABSTRACT

Bottrell, D.G., 1983. The ecological basis of boll weevil (Anthonomus grandis Boheman)
management. Agric. Ecosystems Environ., 10: 247-274.

The boll weevil (Anthonomus grandis Boheman), generally considered to be native to


Mexico or Central America, spread into the southern United States of America in the late
1800s and seriously threatened the cotton industry. As there were no effective alterna·
tives, pest control specialists studied the insect's ecology and advocated cultural practices
that would disrupt its environment and maximize the benefits of natural biological and
environmental controls. An ecologically orientated pest management scheme founded on
cultural practices emerged well before suitable chemical control technology became
available and allowed farmers to live with the weevil problem.
Despite the ingenuity of the early management scheme, it frequently did not provide
satisfactory boll weevil control gauged by present standards. Control of the pest thus
shifted largely from an ecological to a chemical approach as effective synthetic organic
insecticides became available after World War II. The chemical approach was successful
for a number of years, but problems of insecticide-resistant strains of pests, secondary
pest outbreaks, environmental quality, and increased costs of the insecticides have forced
pest control specialists to re-emphasize the nonchemical techniques used widely against
the boll weevil before World War II and to revive the ecological approach to weevil man-
agement.
This article examines boll weevil ecology as related to management of the insect and
reviews the status and prospects of ecologically-based weevil management techniques in
the United States.

INTRODUCTION

The boll weevil (Anthonomus grandis Boheman), generally considered to


be native to Mexico or Central America, was first detected in the United
States in cotton (Gossypium spp.) in southern Texas in the late 1800s
(Howard, 1894). For the next 30 years the insect steadily spread and in-
creased in destructiveness throughout the cotton growing South, but the
challenge occasioned by its adaptation and rapid spread was partially met by
farmers, plant breeders, and entomologists. The longer-season cotton vari-

0167 -8809/83/$03.00 © 1983 Elsevier Science Publishers B.V.


248

eties that had long been noted for high-quality, long-staple fibers were
discarded and replaced by shorter-season, short-staple varieties. The quality
of the new varieties was inferior, but they reached harvest maturity earlier in
the season before the weevil populations caused extensive damage. The
practice of planting the cotton seeds closely spaced in the rows further
hastened maturity and was adopted in parts of the weevil's range. The shift
in cotton from the heavier, more-fertile soils to the lighter, less-fertile soils
in the early 1900s is also traceable to the need to hasten maturity (Klages,
1942).
Some farmers incorporated other elements into their weevil control pro-
gram. Early destruction of and plowing under the postharvest cotton res-
idues eliminated the prehibernating weevils' food source; this reduced the
numbers that overwintered and were thus available to invade and infest the
new plantings in the next season. The farmers also relied on natural biological
and environmental controls.
Though the boll weevil problem was ameliorated, the many excellent
varieties of long-staple cotton were sacrificed, and the farmers who grew the
early-maturing varieties often received lower prices per kilogram of the
harvested lint. Therefore, when synthetic organic insecticides were intro-
duced after World War II, they, plus yield-boosting fertilizers, made it more
profitable for cotton producers to revert to growing the longer-season vari-
eties. The producers could use the new insecticides to protect the cotton
plants from boll weevil damage throughout their extended fruiting period,
thereby permitting greater yields of higher-quality lint, which brought a
better price in the market.
Thus, following World War II, control of the boll weevil, as well as other
cotton insect and mite pests, shifted largely from an ecological to a chemical
approach. Cotton eventually became the crop in the United States most
heavily treated with insecticides, and control of the boll weevil accounted
for nearly one-third of all insecticides used in United States agriculture
(ARS, 1976). The cotton insect control programs based on chemical insec-
ticides were successful for a number of years. However, problems of insec-
ticide resistance, secondary pest outbreaks, environmental quality, and
increased costs of the materials eventually eroded the effectiveness of the
chemical control strategy, forcing the United States cotton industry and the
public research and extension institutions to find alternative solutions.
Important disputes developed in the entomological profession over the
proper approach to pursue, as discussed by Perkins (1983). The current
approach to boll weevil control judged most likely to succeed (NRC, 1981)
is known as 'integrated pest management' (IPM). Integrated pest manage-
ment utilizes a variety of techniques, including, when necessary and appro-
priate, chemical pesticides to reduce pest damage to tolerable levels. First
consideration is given to the use of naturally-occurring mortality elements
of the pest environment, including weather and natural enemies, pest-resis-
tant varieties of crop plants, and cultural practices that destroy the pest's
249

breeding, feeding, or sheltering habitat. The ultimate goal of IPM is to op-


timize pest control in terms of overall economic, social, and environmental
values.
The primary scientific requirement in the development of a successful
IPM scheme for any pest is an understanding of the ecology of the pest and
the cropping system. A reasonable understanding of the actions, reactions,
and interactions of the components of the crop ecosystem affecting the pest
is requisite to an effective IPM program. Only with this knowledge can the
IPM specialist design the optimal pest management strategy and map path-
ways for synchronizing it with and integrating it into the crop production
system(Smith and van den Bosch, 1967).
This paper examines some of the important aspects of boll weevil ecology
as related to the development of IPM programs in the United States.

ORIGIN, TAXONOMY, DISTRIBUTION, AND BIOLOGY OF THE BOLL WEEVIL

Origin, taxonomy, and distribution

Anthonomus grandis was described in 1843 by Swedish taxonomist C.H.


Boheman from a specimen collected at Veracruz, Mexico (Burke, 1968). The
question of the species' original home and original plant hosts has not been
satisfactorily resolved (Burke, 1968), but it is generally considered to be
native to the tropical lowlands of southern Mexico or Central America.
Burke (1968) segregated the species into three taxonomic forms, each
characterized by morphological and behavioral traits and geographic distribu-
tion: southeastern boll weevil (Anthonomus grandis grandis) (Warner, 1966),
which occurs in the southeastern United States, Hispaniola, and parts of
Colombia and Venezuela; thurberia boll weevil (Anthonomus grandis thur-
beriae) (Warner, 1966), which occurs in southern Arizona and northwestern
Mexico; and 'intermediates' or the Mexican boll weevil (Burke, 1968), which
occurs in other parts of Mexico, Central America, and Cuba. As Burke
(1968) and Cross (1973) discussed, the geographic distributions of the three
forms have been complicated by changing crop patterns and by natural and
artificial introductions of the different forms into new areas.
In the United States, the southeastern boll weevil is the primary pest
problem. The thurberia boll weevil in Arizona mostly infests the noneco-
nomic thurberia cotton (Gossypium thurberi Todaro), but is a problem on
commercial cotton in some of its range (Fye, 1968a; Fye and Parencia,
1972).

Reproduction and development

In the United States, the southeastern boll weevil hibernates as a diapaus-


ing adult (Brazzel and Newsom, 1959) in leaf litter or other protected places
outside <the cotton fields. Following emergence from hibernation in the
250

spring, the adults move to young cotton where they feed on leaves and
small squares (floral buds). The adult female usually selects a square ca.
6 mm or more in diameter in which she forms a cavity with her mouth parts,
deposits one egg, and seals with a waxy substance (secreted by her reproduc-
tive organs' accessory glands). The egg stage is followed by three larval
stadia and a pupa, all of which develop inside the square. The effect of the
puncture and early feeding of the larva causes the plant to abscise the square.
The abscised squares generally drop to the soil, although some may remain
on the plant. The emerging adults mate, and the female deposits eggs (up to
400 or more) in the squares; she may also oviposit in small bolls (seed-
bearing capsules). The smaller weevil-infested bolls generally drop to the soil
as do the squares; the larger infested bolls usually remain on the plant and
may continue to grow in an apparently normal manner although the lint in
the infested boll locules may be completely destroyed. The time required for
a generation (from egg to egg) is determined by a variety of factors, but
mostly temperature and growth characteristics of the cotton. The number of
generations per year is therefore variable, ranging from 2 to 8, depending on
the location and environmental conditions (Sterling and Adkisson, 1971).

Host plants

Of the cultivated species of cotton, upland or American cotton (Gossy-


pium hirsutum Linnaeus) is the most significant host of the southeastern boll
weevil. (In the United States, 99% of cultivated cotton is comprised of vari-
eties of this species.) Anthonomus grandis also utilizes Egyptian cotton
(Gossypium barbadense Linnaeus), grown originally in Central and South
America and now chiefly in northern Africa, the Middle East, and the south-
western United States, and several diploid species of Gossypium (D-genome)
in parts of its range (Cross, 1973).
In addition, the boll weevil reproduces on wild malvaceous plants, primar-
ily of five genera: Gossypium, Cienfuegosia, Thespesia, Hampea, and Hibis-
cus (Cross et al., 1975; Bodegas Valera et al., 1977). The insect also utilizes
other plants (wild or ornamental species) as food to sustain its populations
over critical periods, and it may occasionally reproduce on some of these
plants (Cross et al., 1975; Rummel et al., 1978).

ECOLOGICAL RELATIONSHIPS

Weevil-cotton plant relationships

Most cultivated cottons take the form of an annual shrub, 1-3 m tall,
with a vigorous tap root and a dominant ascending main stem which bears,
at each node, a leaf and usually one branch. There are two kinds of branches:
vegetative and fruiting. The vegetative branches are borne at the lower nodes
of the main stem and themselves produce fruiting branches. The fruiting
251

branches, borne at various points beyond mainstem nodes 5-7, are many
jointed. Each joint consists of a short length of stem that terminates in a
floral bud (square) and a leaf, from the axil of which the stem develops
further. The floral buds mature into white or yellow flowers and pollination
occurs on the day the flowers open. The flower petals turn pink and finally
red before abscising, leaving behind the seed capsule known as the fruit or
boll. At maturity, the seed capsules crack and expose the long unicellular
hairs, or lint, borne by the seeds.
That the cotton plant should influence the dynamics of boll weevil
populations seems inevitable. The plant influences the temperature, humid-
ity, and air movement in the insect's microhabitat; fuels growth and repro-
duction; and therefore influences fecundity, development, and survival. In
turn, the weevil influences the plant's development. Abscission of infested
squares and small bolls, triggered by immature weevils developing within,
stimulates floral bud initiation and growth. Low infestations of weevils may
be beneficial when the plants are squaring profusely, but heavy losses of
squares and small bolls affect demand for photosynthate, causing the plants
to grow large and unproductive. (The cotton plant's physiological response
to insect attack was discussed by Gutierrez et al., 1975, 1979,1980.) Heavy
losses of these forms may also restrain growth of weevil populations between
two consecutive generations (Walker, 1966).
Development of eggs in the ovaries is dependent upon the female boll
weevil feeding upon pollen (Fenton, 1952). Before the square flowers,
she drills a hole in it with her long rostrum and feeds on pollen of the
anthers which cover the inside of the square. The opened flowers also furnish
pollen. Adult weevils will also feed on bolls, but without pollen from the
squares or open flowers, reproduction is greatly reduced.
Colonization of posthibernating adults in cotton in the spring is closely
linked to plant phenology (see Ecological relationships of movement).
Colonization progresses rapidly only after the plants are bearing squares
ca. 6 mm in diameter. DeMichele and Bottrell (1976) showed that significant
oviposition does not occur until the squares this size or larger reach a density
of about 25000 ha'", but they did not determine the causative factors.
During the cotton growing season, the growth of weevil populations is
dominated by the physical environment and status of the crop. The size and
abundance of cotton squares are the principal biological factors affecting
boll weevil reproduction and survival (Curry et al., 1980).
The oviposition puncture made by the female weevil does not automatic-
ally interfere with the development of the cotton square. The square contin-
ues to develop normally up to the time of the beginning of the second larval
instar; then, a physiological reaction of the plant to the developing insect
causes an abscission layer to be formed in the peduncle of the infested
square (Coakley et aI., 1969). With the completion of the abscission layer,
the water supply to the square is disrupted and it begins to wilt. The temper-
ature of the square is modified when evaporative cooling of the square bracts
252

and floral leaves (perianth) is severely restricted as a result of the abscission


(DeMichele et al., 1976).
Within a short period after the complete formation of the abscission
layer, the square either falls to the ground or hangs by a fiber thread from
the peduncle. The time from abscission to shedding is probably a function
of crop environment, particularly wind speed and temperature. If the square
falls to the ground, it may be subjected to a variety of microenvironments,
from fully shaded to fully exposed. The particular microenvironment into
which a square falls determines its rate of drying and also the survival of the
immature weevil within it. Survival is a function of the size of the squares,
exposure to sunlight, temperature, and humidity (DeMichele et al., 1976;
Curry et al., 1982). Mortality is highest in insects in small squares (less than
6 mm in diameter) residing in an exposed, hot and dry microenvironment.
This is because the smaller squares dry out faster in this microenvironment
than do the larger squares. The gravid female weevils normally discriminate
against the smaller squares (Jones et al., 1975; Cate et al., 1979), ovipositing
in them only when larger squares are not available. This discriminating be-
havior is obviously important in ensuring survival in the population.
As a member of the insect class, the boll weevil is a poikilothermal organ-
ism. Below a threshold temperature of about 16.5°C there is no develop-
ment. Over the intermediate range of temperatures which the insect normal-
ly experiences in the field, the rate of development and growth is roughly
directly proportional to the ambient temperature above this threshold
(refer to Gilbert et al., 1976, pp. 16-17, for a discussion of the relationship
of insect development and temperature).
At anyone time, the temperature may vary significantly between any two
microenvironments in a given cotton field. Hence, members of a cohort of
boll weevil eggs laid on the same day may show a large variation in develop-
mental time. The food of the weevil larvae may further affect developmental
time. Sterling and Adkisson (1971) found that when both groups were
exposed to the same temperature, boll-reared insects required 11-31 days
more than square-reared insects to reach the adult stage. The larval food may
also affect diapause in the adult stage. Sterling (1971) showed that the per-
centage of diapause is higher in adult boll weevils emerging from bolls than
from squares. Survival during hot and dry periods is also probably higher in
boll-reared than in square-reared weevils because the infested bolls are not
as likely to abscise as the infested squares are and fall to the hostile ground
environment (Sterling and Adkisson, 1978).
Of course, the variability in developmental time and in other biological
parameters has a genetic basis. Variability would occur even if a cohort of
weevils were reared under identical microenvironments. However, for many
biological parameters there has been limited work to determine the extent
to which the observed differences between the phenotypes of different
weevils are due to heredity and the extent to which they are environmentally
induced.
253

Relevance to boll weevil management

Cotton has evolved diverse means to avoid, tolerate, or recover from


attacks of the boll weevil (Gutierrez et al., 1979). Plant breeders have used
this natural process to their advantage by deliberate selection of agronom-
ically acceptable cotton varieties that withstand the pest's attack. Develop-
ment and use of such varieties is a proven, effective, economical, and safe
method of boll weevil control, ideally suited for integrated pest manage-
ment, and is compatible with other desirable pest suppression methods such
as biological control.
As the boll weevil spread in the southern United States towards the turn
of this century, efforts were made to propagate races of cotton which
resisted its attacks. Special interest attaches to the investigations made by
Cook (1904a) in Guatemala. Examination of cotton being propagated by
Guatemalan Indians showed that although the weevils attacked young
squares, these forms did not drop off because of a special growth of tissue
inside which frequently killed the eggs and young larvae. In addition, inside
the young bolls which had been pierced there was a similar proliferation or
growth of the tissue which crushed and killed the insects. As a result of long
contact between the cotton plant and the boll weevil, and probably assisted
by the unconscious human selection of the seeds of the healthier surviving
plants for use in the next plantings over long ages, this cotton type evolved
in Guatemala (Cook, 1904a). These discoveries encouraged plant breeders
in the United States to exploit boll weevil resistance in cotton breeding
programs.
Cotton breeders soon discovered that early-maturing varieties of cotton
were more apt to escape boll weevil damage. Therefore, as noted in the
Introduction, the longer-season, long-staple cotton varieties that had long
been used for high quality were discarded and replaced by shorter-season,
short-staple varieties. Use of the early-maturing varieties made it possible for
the United States cotton industry to survive the boll weevil plague of the
early 1900s (Helms, 1980).
The search for boll weevil resistance in cotton germ plasm and transfer of
resistance into commercial varieties has attained great momentum in the past
few years (Jones et al., 1978; Benedict and George, 1979). Several genetic
sources of resistance have been found, and they are being utilized in cotton
breeding programs. The development of agronomically acceptable early-
maturing varieties has received considerable emphasis.
The shorter-season varieties escape much of the buildup of prehibernating
boll weevils (Walker and Niles, 1971) and also other late season insect pests
(e.g., pink bollworm, Pectinophora gossypiella (Saunders)). A second advan-
tage is that they may be harvested before the occurrence of fall rains that
frequently damage cotton quality, reduce yield, and prevent the early
shredding and plowing under of cotton stalks for control of the weevil and
pink bollworm (Adkisson et al., 1982).
254

New techniques in genetic engineering offer potentially great payoffs in


developing cotton plants that resist, tolerate, or escape weevil attack. How-
ever, it will probably be many years before these techniques can be tried
out.
The cotton plant has a limited capacity to set bolls, depending upon
growing conditions and individual plant health. Cotton fruits in excess of
this capacity will drop from the plant in absence of boll weevils or other
fruit-feeding pests. Therefore, insect-caused injury of the surplus fruits
cannot be included in economic damage, and it is uneconomical to apply
controls that protect these fruits (Smith and van den Bosch, 1967). On the
other hand, when insect feeding reduces a boll load to a level below the
cotton plant potential, there is a crop loss.
The decision, then, to attempt to reduce the boll weevil population
depends upon the value of the crop that can be recovered or protected,
balanced against the costs (economic, environmental, and social) of the
control measure. Making an intelligent decision requires information on the
'economic threshold'. The economic threshold is the population level at
which a pest (boll weevil) has just attained 'real' pest status. In other words,
it is the density of the population below which the cost of applying control
measures exceeds the losses caused by the pest (Stern, 1973). Understanding
the weevil-cotton plant relationships is requisite to establishment of eco-
nomic thresholds used in integrated pest management programs.

Weevil-wild plant relationships

Most of the wild plants known to host the boll weevil occur in Latin
America. With a few exceptions, host plants other than cultivated cotton
are relatively unimportant in the United States (Cross et al., 1975). The most
important wild plants in the United States are Cienfuegosia drummondii
(A. Gray) Lewton, a host of the southeastern boll weevil in the Costal Bend
region of Texas, and thurberia cotton, a host of the thurberia boll weevil
in Arizona, as discussed above.
The weevil readily develops on Cienfuegosia plants in natural situations
even when the plants occur several km from cotton. In some areas of Texas,
there is no evidence that any regular exchange of weevils takes place between
Cienfuegosia and cotton, whereas in other areas there is some exchange be-
tween the two host plants. Burke and Clark (1976) hypothesized that the
weevil probably could maintain small populations on this wild host in the
absence of cultivated cotton.
In Arizona, there apparently is minimal exchange of the thurberia boll
weevil between its normal thurberia cotton host and cultivated cotton
except where thurberia cotton grows adjacent to cultivated cotton (Fye and
Parencia, 1972).
In the tropics of Chiapas, Mexico, the Mexican boll weevil readily feeds
and reproduces on Hibiscus tiliaceus Linnaeus (Bodegas Valera et al., 1977).
255

This mallow occurs in many places where Anthonomus grandis occurs, but
the insect has been found to utilize the plant only in Chiapas. Bodegas Valera
et al. (1977) postulated that weevils in Chiapas freely exchange between
Hibiscus tiliaceus and cotton growing nearby. However, there is only limited
quantitative information on the interchange of boll weevils between this wild
host, or any other wild host, and cultivated cotton.

Relevance to boll weevil management

Burke (1968), among others, stressed the importance of ecological studies


of the boll weevil on its wild host plants in the American tropics. Without
these studies, there remains a major void in knowledge of the genetics,
systematics, ecology, and biogeography of the weevil species.
Burke and Cate (1979) recently described a new species, Anthonomus
hunteri Burke and Cate, the closest known relative of the boll weevil, repro-
ducing on the wild plants Hampea nutricia Fryxell and Hampea trilobata
Standley in the tropics of southern Mexico. Before this species was discov-
ered, no close relative of the boll weevil was known, and without this tax-
onomic link, efforts to pinpoint the boll weevil's place of origin, original
hosts, and pathways of dispersal were hampered. Discovery of the close
relative therefore has more than academic significance. It may be one step
closer to resolving questions concerning the boll weevil's place of origin
and original hosts. Studies of the insect in such areas where there has been a
very long insect-plant association would greatly widen the ecological
understanding of the species and may lead to important sources of more
effective natural enemies and host plant resistance factors.

Community relationships of hibernating weevils

As the boll weevil spread in the United States, it entered and adapted to
temperate climates which differed drastically from the hot tropical climates
in its southern range. The species now occurs at northern latitudes above
36° where winter temperature occasionally drops below -20°C and the
autumn-to-spring frost period spans about four months. To survive the
winter, adult weevils of the southeastern form seek protected hibernating
shelter outside the cotton fields. (The thurberia weevils spend the winter as
diapausing adults in cells in bolls of their wild host plants.)
A combination of extended cold temperature (O°C or lower) and very dry
conditions will reduce survival in the hibernating population (Pfrimrner and
Merkl, 1981). Other weather factors probably also affect winter survival,
and combinations of factors may interact to magnify or minimize the tem-
perature or other specific effects. For example, accumulation of snow during
extremely cold weather may serve as insulation against the cold (Pfrimmer
and Merkl, 1981). Whether boll weevils survive the winter weather greatly
depends on the type of hibernating habitat.
256

The early boll weevil literature frequently referred to wooded areas in


river bottom lands as 'ideal' hibernating habitat and stated that bottom-land
cotton near such areas was much more likely to become infested with wee-
vils earlier in the season than cotton in fully open terrain. Shedding of leaves
of deciduous broadleaf trees and shrubs in bottom lands results in the
accumulation of deep litter (leaves, stems, and duff) under which the boll
weevils crawl. The thick litter provides insulation from the cold air and
retains moisture during dry periods of the winter.
The influence of hibernating habitat is particularly apparent in areas
where habitat cover is sparse, for example, in northwestern Texas. In this
area of Texas, only ca. 25% of the cotton fields abut on what is considered
to be a primary hibernating habitat for the weevil (by contrast, 100% of the
cotton fields in some areas of the United States may be surrounded by
primary hibernating habitat). Whether cotton fields in northwestern Texas
become infested with posthibernating boll weevils depends largely on the
type of hibernating habitat in the vicinity and the distance of the cotton
fields from the habitat (Rummel and Adkisson, 1970). Most of these cotton
fields that become infested with the posthibernating populations are located
close to densely vegetated native plant communities (0.15 km or less).
Searches for boll weevils during winter have produced large numbers hiber-
nating in these same native communities and shelterbelt plantings (Bottrell
et al., 1972; Slosser and Boring, 1980).

Relevance to boll weevil management

Some of the early references to control of the boll weevil advocated prac-
tices that would suppress the pest in or near hibernating habitat. Isely (1930)
demonstrated the value of destroying vegetation along the cotton field
edges, in fence rows, and ditches where weevils hibernate.
Isely (1926) also demonstrated the value of restricting the applications of
insecticide (calcium arsenate dust) to those areas of the cotton fields where
the posthibernating weevils colonized in the spring after emergence from
hibernation. These restricted treatments to cotton, made in the loci of the
hibernating habitat, effectively controlled the weevil pests while sparing
natural enemies in the rest of the field, thus minimizing both costa of appli-
cation and adverse effects on nontarget organisms.
The basic technique developed by Isely (1926) is currently used quite
effectively against the boll weevil in parts of the United States. In t'e spring,
farmers or their field scouts routinely check the cotton fields for Posthiber-
nating weevils developing near the hibernating habitat and selectively treat
the 'hotspot' infestations. In Texas, pheromone traps are being utilized as
an aid to detect weevil movement in these hotspot areas, (see Ecological
relationships of movement).
Slosser and Boring (1980) listed various methods for controlling the wee-
vils within the hibernating habitat - use of herbicides, mechanical means,
257

and fire to destroy the hibernating quarters; use of insecticides and biological
agents; grazing of cattle, etc. Each method has certain advantages as well as
limitations. Massive destruction, chemical alteration, or any other modifica-
tion of the vegetation within the habitats may be ecologically disruptive.
Therefore, the potential benefits of approaches aimed at the hibernating
populations should be carefully examined along with the potential costs to
the environment.
These approaches would probably have limited value in an area with mild
winters such as the Lower Rio Grande Valley of Texas where boll weevils
utilize a much wider range of hibernating habitat (Graham et aI., 1978) than
they do in northwestern Texas, for example.

Natural enemy-weevil relationship

After the boll weevil entered the United States toward the turn of this
century, entomologists took steps to ascertain and encourage its biological
control by natural enemies. The first records of natural enemies were made
by Townsend (1895), when he mentioned both a hymenopterous (wasp)
parasite and the suspicious occurrence of predatory insects (syrphid fly
larvae and coccinellid adults) in the squares. In Guatemala, Cook (1904b)
found that the weevil was eaten by the 'kelep' or Guatemalan ant (Ecta-
tomma tubercula tum Olivier). He introduced colonies of the ant in Texas,
but they did not survive the cold winters.
In 1912 Pierce published "The Insect Enemies of the Cotton Boll Weevil",
which is still the most thorough and stimulating analysis of biological control
of the pest. Pierce listed 49 species of insects and mites (29 parasites and 20
predators) known to attack immature stages of the pest. He reported that
actions of the natural enemies accounted for more than one-third of all
mortality occurring in immature weevils. In addition, he listed six species
of predatory insects which preyed on adult weevils.
Pierce (1912) emphasized the importance of relationships of weevil
enemies with the other insects (co-hosts) utilized by the enemies, and
he proposed ways to increase biological control of the weevil by manipulat-
ing these co-hosts. By destroying ragweeds (Ambrosia sp.), situated next to
a cotton field and infested with Lixus beetles - hosts of boll weevil para-
sites (Eurytoma sp.), he increased the level of parasitization in pests in the
cotton. Destruction of the weeds eliminated the Lixus infestations; then,
apparently the parasites which were left behind shifted to the weevil pests
in the cotton field. The long-term effect of this practice was not determined,
however.
Others working on the pest in the early 1900s emphasized the importance
of biological control. However, work in this area dwindled in the early 1920s
and essentially stopped in the late 1940s when synthetic organic insecticides
became available. This work has received only limited emphasis in recent
years.
258
OOOOC'l 259
10';<

OOOC'l~O
'<ji""';C'iaiai
...... Ol

OOOt-~
In 1971, Cross and Chesnut listed 42 species of insects and mites known
to parasitize the species throughout its range. More recently, Cate and Clark
(1978) and de Coss Flores et al. (1977) discovered several new parasitic
species and found about 11 insect parasite species attacking the boll weevil
(or the closely related species Anthonomus hunteri) in cotton and wild host
plants in Mexico. The most common insect parasite of boll weevil in the
United States is the wasp Bracon mellitor Say, whose larvae feed on the
pest's larvae and pupae.
A variety of invertebrate and vertebrate predators attack the boll weevil.
The ant Solenopsis invicta Buren is apparently one of the most effective
predators (Sterling, 1978). Under certain conditions, birds may be impor-
tant. Howell (1907), for example, reported 43 bird species feeding on the
weevil and proposed that legislation be enacted to protect them.
Few pathogens have been found attacking the boll weevil in nature
(Cross, 1973). However, surveys for pathogens have been limited and have
rarely encompassed the American tropics.
Quantitative evaluations show that where insecticides are rarely used, the
native insect parasite Bracon mellitor may cause high weevil mortality at
certain times. In Texas cotton not treated with insecticides, the parasite was
shown to exhibit a weak density-dependent response to its weevil host, that
is, Bracon-inflicted mortality increased proportionately with weevil density
(Bottrell, 1976). The effect of the parasite on its weevil host is shown in
Table I. Although significant at times, Bracon mellitor is insignificant at
other times, and it induces high levels of mortality only after the weevil has
caused appreciable damage to the cotton. It attacks hosts other than boll
weevil and therefore is not dependent on the pest for its existence.
Parasites possessing high host specificity or capacity to develop high
preference for the target species are generally the best candidates for biolog-
ical control (Huffaker et al., 1977). A generalist parasite such as Bracon mel-
litor would not be expected to provide fully effective economic control of
the boll weevil in most situations.
Although natural enemies do not appear by themselves to provide reliable
economic control of the weevil under eixsting cotton production conditions
in the United States, they may have considerable potential in integrated
pest management. As Curry et al. (1980) pointed out, any tactic such as
biological control when used unilaterally, or naturally occurring as the only
mortality factor, must induce very high levels of weevil mortality to be
dependable. On the other hand, a tactic may provide only minimal levels of
mortality and still contribute significantly when used alongside other tactics
in IPM systems.

Relevance to boll weevil management

Classical biological control, involving deliberate introduction and estab-


lishment of natural enemies in areas where they did not previously occur, has
been highly successful against some insect pests of foreign origin, leading
some to believe that this approach may have promise against the boll weevil
in the United States. However, efforts to establish exotic natural enemies
(from. Mexico and Central America) of the weevil have been unsuccessful.
The wasp parasite Heterolaccus (= Pteromalus) grandis Burks, procured in
Mexico, significantly reduced weevils in cotton when released in Mississippi
during the summer (Johnson et al., 1973), but did not survive the winter in
Mississippi, presumably a victim of the cold weather. Possible alternative
approaches would be to rear large numbers of the non-adapted parasite for
periodic release during the summer, to select for increased cold tolerance,
or to seek strains from colder (mountainous) parts of Latin America.
McGovern and Cross (1976) found that Bracon mellitor caused signifi-
cantly higher mortality in boll weevil populations in frego-bract cotton
varieties than in normal-bract varieties. The floral bud bracts (leaf-like
structures on the floral buds) of the frego-bract varieties are rolled up, a
condition which exposes the floral buds. By contrast, the floral bud bracts
of the normal bract cotton tightly enclose the floral buds. The female
parasite's success in searching for and finding immature weevils inside
infested floral buds is greater in cotton with the exposed frego-bract floral
buds.
Cotton varieties possessing the frego-bract character (Lincoln and Waddle,
1966) have a high level of genetic resistance to the boll weevil (Jenkins and
Parrott, 1971). Yet, it is unlikely that this resistance is sufficient by itself to
keep weevils from causing economic damage. However, IPM schemes based
on the use of frego-bract cotton in combination with practices that protect
and increase parasite populations may have promise in some areas.

Ecological relationships of diapause

The adult boll weevil enters a state of facultative diapause (Brazzel and
Newsom, 1959). This state results in survival of populations during periods
of host plant unavailability and environmental adversity.
To survive the winter in the colder areas of the United States, the south-
eastern boll weevil enters a state of 'hibernal' diapause and the adults dis-
perse from cotton to leaf litter and other protected environments outside
the cotton fields. In the hibernal diapausing state, the insect's reproductive
organs do not function, excess body fat accumulates, and water content of
the body is very low. In southern Texas, where the winters are milder, the
adult weevils remain active during the winter months and do not exhibit
the morphological and physiological characteristics that diapausing adults
in colder areas do (Guerra et al., 1982). In Arizona, the thurberia boll
weevils diapause as unfed adults entrapped in the larval cells in capsules of
their wild host plant, thurberia cotton (Fye, 1968a; Fye and Leggett, 1969).
In the tropics of southern Mexico and Central America, the Mexican boll
weevil spends the dry season (cotton-free period) encapsulated in the cotton
bolls in a state of diapause (Flores Garda et al., 1977), known as 'aestival'
or 'swnmer' diapause. Weevils of the southeastern form may also enter a
state of aestival diapause (Phillips, 1976), but in this state the adult weevils
are not confined in the cotton bolls as the diapausers in southern Mexico and
Central America are; some may even feed.
A variety of environmental factors, including photoperiod, temperature,
and quantity and quality of food, are suspected to induce diapause in the
boll weevil, as discussed by Cross (1973). However, all the interrelated
factors that may affect diapause have not been assembled into a coherent
whole.
In the United States, diapausing weevils (southeastern form) begin to
appear in the population when cotton plants cease flowering and approach
maturity. Studies in Arkansas suggest that any condition which forces the
cotton plant into a state of stress may induce diapause; populations devel-
oping on stressed (dry, non-flowering) cotton during the summer may enter
aestival diapause several weeks prior to the usual hibernal diapause period
(Phillips, 1976). Generally, in the southeastern form, the portion of dia-
pausing members in the population increases steadily from mid- to late-sum-
mer until frost. However, not all late-season adults exhibit diapause; some
reproducing weevils inhabit the cotton fields at the time of frost.
In the colder areas of the United States, the winter diapause state will
persist in some individuals until host plants capable of supporting boll weevil
reproduction become available in the spring. Gaines (1959), among others,
noted that weevils may become active and fly during normal hibernation if
temperatures exceed about 16.5°C. However, it has not been determined if
such individuals are truly in diapause or if they are capable of surviving
until host plants become available in the spring.

Relevance to boll weevil management

Entomologists working on the boll weevil in the United States towards the
turn of this century recognized that early harvest and destruction of and
plowing under the harvested cotton plants would reduce prehibernating
populations of the pest (Townsend, 1895; Mally, 1901). These practices,
though ecologically sound and potentially effective, were not widely used
(Rummel and Frisbie, 1978). One obstacle to widespread adoption related to
the labor requirement; hand harvesting the crop was highly labor demanding
and slow. Another obstacle related to the equipment that farmers used to
destroy and plow down the postharvest plants; the equipment was slow
and inefficient gauged by present standards.
The report by Brazzel and Newsom in 1959 that boll weevils enter a state
of diapause revived interest in the development of control approaches aimed
at the prehibernating populations,
The most widely adopted weevil management program which is designed
around the phenomenon of diapause is the 'diapause control' approach
developed by Brazzel (1959). This approach, as opposed to the conventional
insecticidal control programs directed against reproducing weevils during the
primary cotton growing season, utilizes insecticides near the cotton harvest
period to kill the diapausing insects. (If applied early enough in the season,
the insecticides also kill large numbers of weevils of the season's last repro-
ducing generation, thus suppressing the reproducing females before they
deposit eggs which would develop into diapausing adults.) Applications of
the insecticides correspond to a period when diapausing adults begin to
appear in significant numbers, but before they fly from the cotton fields to
hibernate. Knowledge of the seasonal occurrence of diapause in the weevils
and of the time of their hibernating flights is requisite to proper timing of
the applications. If the numbers of diapausing weevils are sufficiently re-
duced (90-99%), the development of damaging infestations may not occur
until late in the subsequent cotton growing season (Adkisson et al., 1966).
Experience has shown that all cotton farmers in the area must cooperate in
the diapause control program or it may not be effective. As discussed later,
the boll weevil is a strong flier. A farmer's efforts therefore can be cancelled
by weevils flying from a neighbor's field not receiving the diapause control
treatment.
The diapause control method described by Brazzel (1959) and modifica-
tions thereof (Adkisson et al., 1966; Lloyd et al., 1966; Rummel et al.,
1975a) have been used successfully in several cotton growing areas of the
United States. By applying the insecticides towards the end of the cotton
growing season, the method achieves ecological selectivity in the chemicals:
reduction in insecticide use during the regular growing season results in less
harm to natural enemies important in the cotton ecosystem and surrounding
areas. The late-season applications often eliminate the need to control first
and second, or even later generations of boll weevils in the following season,
thereby reducing the number of insecticide applications 40-50% compared
to conventional treatment with insecticides early in the season.
The benefits of diapause boll weevil control have been impressive in some
areas, as illustrated by a program in northwestern Texas (Rummel et al.,
1975a). During the period 1964-1973, a total of 4.13 million kg of insecti-
cides (nearly all malathion) was applied to cotton (47773-145881 ha
treated each year) at a total cost of US $9.8 million (1974 value). Econ-
omists (Lacewell et al., 1974) calculated that the program saved cotton pro-
ducers in the affected area 3.74-9.28 million kg of insecticide and US
$12.7-21.3 million in production costs annually, and it increased the cotton
yield 4-7%.
Diapause boll weevil control is not a panacea, however. The possibility
that weevils will evolve strains resistant to the organophosphorus insecticides
used in the control programs is one potential drawback. In the diapause
control program in northwestern Texas, noted above, prehibernating adults
have been subjected to heavy insecticide pressure for more than a decade,
but studies have not revealed any genetic resistance to the insecticides
(Pruitt et al., 1978). However, with continual insecticidal pressure on the
population, it is almost certain that resistance would eventually develop.

Ecological relationships of movement

It is by flight that the boll weevil has dispersed to wide-ranging environ-


ments of tropical rainy, dry, and humid mesothermal climates. The spread
of the insect in the southern United States at the tum of this century illus-
trates its remarkable powers of dispersal and adaptation.
In the spring, the southeastern boll weevils emerge from hibernating sites
outside the cotton fields and disperse. Emergence usually extends over several
weeks. Wade and Rummel (1978) found a positive relationship in the time
of entry of diapausing boll weevils into hibernation in autumn and the time
of their exodus from hibernation in spring; the first insects to enter hiberna-
tion were the first to emerge from it, and so on. Mortality may be quite high
in the posthibernating population. White and Rummel (1978) estimated that
only 5-10% or less of the population emerging from hibernation survive to
colonize cotton in northwestern Texas. 'Suicidal emergence', that is, spring
emergence from hibernation and death prior to the availability of host
plants, is one important mortality factor. Other mortality factors (predation,
storms, etc.) may be involved, but there are only limited data on the mortal-
ity in posthibernating adult populations.
The maximum distance that the posthibernating insects can fly in search
of cotton is not known. They will sometimes infest cotton several km from
their hibernating sites; however, Rummel and Adkisson (1970) showed that
most posthibernating weevils colonized cotton fields situated very near
(0.15 km or less) the hibernating sites.
Hardee et al. (1969) hypothesized that cotton plant odors play little,
if any, role in triggering colonization of the posthibernating weevils. They
argued that the first individuals to colonize cotton find it by chance, but,
after feeding on the cotton plants, males produce an aggregating pheromone
which attracts both females and males, and this is the primary factor in-
volved in weevil colonization. However, White and Rummel (1978) dem-
onstrated the important role of plant phenology in colonizing posthibernating
weevils, and their findings agreed with those of Walker and Bottrell (1970),
Roach et al. (1971), and Rummel and Bottrell (1976). Very few of these
insects enter the cotton before the plants support squares that are ca. 6 mm
in diameter. Colonization then progresses rapidly. The aggregating phero-
mone emitted by male weevils that have colonized and fed on the squares is
a powerful attractant and undoubtedly speeds up colonization, but the
squaring plants themselves appear to trigger colonization of the first major
group of recruits in the cotton. Plant odors are probably involved, but the
discriminating chemical sense of the boll weevil is not well known.
For several weeks following colonization of the posthibernating adults,
movement is periodic, characterized by crawling and short flights within
or near cotton fields and generally not far beyond them (Hunter and Hinds,
1905). Isely (1926) reported that each period of in-field dispersal coincides
with the emergence of new-generation adults. The spread of adults across a
field is usually direct from plant-to-plant or row-to-row and not the result of
long flights. Gaines (1932) observed a slow, progressive seasonal spread of
boll weevils outward from the loci where the posthibernating adults colo-
nized. In 9 weeks the infestations had spread about 1.2 km, and in 12 weeks
they had spread about twice that distance.
Movements occurring in the later generations involve spectacular exodus
flights from the cotton fields in all directions. These annual flights begin
approximately the same time each year in a given location. Hunter and Hinds
(1905) were among the first to report on these spectacular occurrences:
"One of the most striking facts observed was that the weevils succeeded in
crossing bodies of water which, in some cases, were fully 10 miles (16.1 km)
in breadth. Stretches of non-cotton-producing country were also passed
which were fully 40 miles (64.4 km) in breadth. During the entire period of
migration it was noticeable that weevils in the fields took wing far more
readily than they had done earlier in the season. In many cases, upon slight
disturbance, weevils were seen to rise above the roofs of houses and the
tops of tall trees and disappear from view. How high they ultimately rose in
flight no one can say, for the eye could not follow them farther".
During the annual late-season flights, boll weevils have been recorded to
cross open water up to 16.1 km wide (Hunter and Hinds, 1905); to infest
cotton 41.2-54.7 km (Beckham and Morgan, 1960) and 64.4 km (Hunter
and Hinds, 1905) from a known source; to move more than 225 km in three
days when helped by wind (Hinds, 1916); and to respond to traps baited
with male boll weevils or synthetic boll weevil pheromone at distances of
72 km (Davich et al., 1970) and 144 km (NRC, 1981) from a known source.
Weevils marked and released have been recovered at distances of up to ca.
72 km (Johnson et al., 1976).
The distance that the weevils travel during the late season is probably
related to many factors; certainly, the meteorological situation affects the
airborne movement. Taft and Jernigan (1964) and Rummel et al. (1977)
showed that those weevils dispersing during late season fly at considerably
greater heights than ones dispersing during the spring. Fye (1968b) reported
that air turbulences (thermal convection and 'dust devils' or whirlwinds)
occurring in late summer in Arizona and neighboring Mexico probably
elevate the weevils into upper wind currents. He postulated that these
turbulences could move them in the direction of the air flow and therefore
account for long-range transport.
The factors that evoke the annual late-season flights are not known, al-
though some have implied that the phenomenon is related to the develop-
ment of dense populations of weevils and the cessation of fruiting by the
cotton plant. Harned (1910) stated that the insects spread slowly when food
is plentiful and rapidly when it is scarce. However, Hunter (1909) and
Hunter and Coad (1923) recognized that the 'instinct to migrate' by weevils
in late season is not necessarily associated with too little food for the popula-
tion. Bottrell and Moody (1973) reported that the beginning of late-season
movement in a given area showed a similar recurring pattern each year, al-
though weevil density and cotton phenology varied from year to year. In-
deed, the association of this movement with large populations and various
other factors may be merely correlative rather than one of cause and effect.
However, the quality of the food obviously affects reproduction and ovi-
position: if plentiful food hastens ovary development and lack of it length-
ens the preoviposition period, food may also affect migration directly (John-
son, 1969).
To determine the reproductive status of boll weevils dispersing during late
season in northwestern Texas, Rummel et aI. (1975b) regularly monitored a
small cotton plot (6.1 X 8.1 m) isolated by about 4.8 km from weevil-
infested cotton. Of the 82 weevils (46 females, 36 males) that entered the
plot during the period August 5-November 11, 86% were classified as
reproducing and 14% as diapausing. Most (71%) of the diapausing insects
were recorded after October 24. Wade and Rummel (1978) showed for this
area of Texas that most diapausing weevils enter hibernation in October and
November.
Rummel et al. (1975b) and Walker and Bottrell (1970) showed that the
annual late-season flights are initiated by reproducing insects which fly to
new breeding areas; the portion of reproducing insects in the population
then decreases as the season progresses and diapausing insects begin to fly
in search of hibernating sites. Most diapausing weevils fly only short dis-
tances from cotton fields before entering hibernation (Fye et aI., 1959;
Beckham, 1957; Bottrell et aI., 1972). Fye et al. (1959) discovered that
90% of the weevils hibernating in woods surrounding cotton fields in South
Carolina occurred within 290 m of those fields.

Relevance to boll weevil management

Townsend (1895) was the first in the United States to recognize the
significance of boll weevil movement when developing policies and programs
to combat the pest. Soon after the discovery of the weevil in southern Texas,
he outlined extensional, legislative, and farmer organizational needs for its
control. He recognized that the weevil possessed remarkable powers of
dispersal and would spread from Texas to other cotton states if drastic
measures were not taken. He therefore recommended that a cotton-free zone
be maintained north of the infested area in Texas so as to counter spread of
the insect into other areas of the United States. His recommendation was not
followed, however, and the weevil continued to spread.
The current limited understanding of many aspects of boll weevil move-
ment hinders progress in developing ecologically-based management systems
for the pest. Examples are given to illustrate the value of knowledge on
movement in the development of these systems.
The diapause control approach (Brazzel, 1959), discussed above, is in
effect a strategy aimed at restraining late-season boll weevil movement. The
objective of the late-season applications of insecticide is to reduce the
numbers of diapausing weevils that emigrate to hibernating sites; if the
applications are timed early enough, they also restrain the flights of the last
generation of reproducing weevils capable of emigrating many km and
invading new territories. This approach has successfully prevented the
spread of weevils from infested cotton in northwestern Texas to uninfested
cotton in neighboring areas of Texas and New Mexico (Rummel et al.,
1975a). An understanding of the chronological onset of diapause and the
occurrence of late-season flights in the weevil population is necessary to
determine when to apply and relax the control procedures so as to achieve
the most effective results (Wade and Rummel, 1978).
The use of boll weevil traps baited with synthetic pheromone has received
extensive publicity in recent years. Traps often produce spectacular catches,
leading some farmers and pest management specialists to believe that the
traps are having a major impact on the pest population, but the catches in
themselves are not proof of the traps' effectiveness unless they can be related
to a decreased pest population in the area.
In northwestern Texas, White and Rummel (1978) found that only the
last 10% or less of the boll weevil population to emerge from hibernation in
the spring colonizes cotton. Therefore, pheromone-trap destruction of the
first 90% of the emerging population may have no beneficial impact in re-
ducing the severity of boll weevil damage. Without such knowledge as this on
the relationship between spring movement and colonization in cotton, the
pheromone-trap catches in themselves could greatly misrepresent the traps'
true value (Ridgway et al., 1976).
Pheromone traps do have considerable value in population monitoring,
and trap data have been used to aid in determining the need to control post-
hibernating boll weevils colonizing in cotton. In Texas, the traps (6-8 per
cotton field 20-121 ha in size) are evenly spaced around the field's margin,
and are examined routinely for weevil catches during the period that the
posthibernating adults colonize cotton. The decision to apply insecticide
is based on the 'trap index' system developed by Rummel et al. (1980). The
size of the trap index determines if control is required. The trap index
system is more accurate than field scouting is in determining the need for
control (Rummel et aI., 1980).

PROGRESS WITH HOLISTIC IPM SYSTEMS

Many cotton farmers in the United States where the boll weevil is a
problem are now using IPM techniques. A substantial portion of the United
States cotton land is being regularly monitored by private, college-trained
pest management consultants. These consultants offer a wide range of farm
services, including soil analysis, fertilizer recommendations, crop variety
selection, and advice on infestation of boll weevils and other pests, pesticide
application, and alternative control methods. The extension services are also
engaged in promoting these activites. This approach facilitates more efficient
use of pesticides than is possible without the field checking and the other
services, and in most areas it has meant a significant reduction in pesticide
use without a loss in yield or quality. In Texas, where one-half of the cotton
in the United States is grown, insecticide use on the crop dropped from 8.75
million kg in 1964 to 1.04 million kg in 1976 (OTA, 1979), and use of IPM
techniques by farmers undoubtedly contributed to this reduction (Adkis-
son et al., 1982).
However, there are few true, comprehensive IPM programs now in use in
boll weevil infested cotton. Most of them have been unilaterally developed
for the weevil or this insect pest and a few other cotton insect pests that
occur sympatrically with it, and not whole pest complexes (Phillips et al.,
1980).
Farmers are rarely confronted by a single pest problem, but rather, by
complexes of pests: different kinds of insects; mites; disease-causing organ-
isms; weeds; and sometimes other pests. Measures to control the boll weevil
directly or employment of a desirable agronomic practice that reduces its
severity may create or intensify other pest problems which may also be
multiplied by extremes in weather. It is obvious, then, that pest control
recommendations cannot be evolved independently for weevils; nor can
optimal control strategies be developed without considering the cotton
production system as a whole. Multipest integrated management schemes,
synchronized with and integrated into optimal cotton production systems,
are essential for long-term profits. Efforts to develop such multipest IPM
schemes have just begun (OTA, 1979).
CONCLUSIONS
The genesis of ecologically orientated boll weevil management traces back
to the late nineteenth and early twentieth centuries. Particularly outstanding
contributions of this period include the work of Mally (1901), Cook (1904a),
Hunter and Hinds (1905), Pierce (1912), Hunter and Coad (1923), and
Isely (1926). As there were no effective alternatives, these specialists studied
the weevil's ecology and advocated cultural practices that would disrupt the
pest's environment and maximize the benefits of natural biological and
environmental controls. From these efforts, some of the most ingenious
pest management systems ever developed for agricultural insect pests evolved
(Smith et al., 1976; Bottrell and Adkisson, 1977).
However, emergence of the synthetic organic insecticides halted the
ecological approach to boll weevil management. As Fye (1974) pointed out,
the 'sledge-hammer' approach of insecticidal control led to less sophistica-
tion, less complexity, and less need for ecological analysis.
Problems experienced with the unilateral chemical control strategy
renewed interest in ecologically orientated boll weevil management. The
revival of this ecological approach offers special opportunities for modern
crop protectionists. The science of ecology has greatly advanced in recent
decades; hence, modern specialists may draw on many skills that were not avail-
able to their early twentieth century predecessors (Huffaker, 1972). Armed
with systems analysis, mathematical models, and computers, the modern
specialists are better equipped to analyze complex ecological systems. The
model of Curry et al. (1980), the most comprehensive boll weevil model to
date, is giving a clearer understanding of the various interactions in the boll
weevil's life system and the cotton ecosystem, and is leading to new path-
ways for achieving optimal integration of IPM tactics. However, computer
models are not a panacea, anymore than chemical pesticides are, and as
ecological tools they are no stronger than the ecologists using them.
The development of comprehensive, multipest IPM schemes for weevil-
infested cotton will be a long-term effort requiring boosted collaborative
teamwork of pest control specialists, ecologists, plant breeders, soil scientists,
economists, sociologists, system scientists, and others. Basic biologists -
taxonomists, geneticists, etc. - must be integral to these teams. The ecol-
ogists must take a primary lead, but a special kind of ecologist is needed. At
present, too few ecologists are able to bridge the gap between pest control
practitioners who are blinded by short-term needs and theoreticians who
have inadequate appreciation of practical realities (Way, 1973).
A variety of technical, economic, and attitudinal barriers slow progress in
the development and implementation of ecologically orientated boll weevil
management in the United States. Perhaps the single largest obstacle in
recent years can be traced to the United States Department of Agriculture's
(USDA) involvement in a scheme to eradicate the boll weevil from the
United States, discussed by Perkins (1983). The USDA, the major federal
institution involved in pest control research, has invested millions of dollars
in research attempts to develop various techniques which, when integrated
together, will achieve complete annihilation of the boll weevil from the
United States. Research to use these techniques in IPM schemes aimed only
at keeping the boll weevil below damaging levels has suffered as a result.
Furthermore, fundamental research on biological control, cultural control,
and ecology - the backbone components of IPM - has been neglected
because the USDA has not viewed this research as essential to an eradication
effort. Yet, as discussed by Perkins (1983), all of the scientific committees
outside of the USDA that have evaluated the institution's program in erad-
ication research and implementation have seriously questioned whether
eradication of the pest from the United States is a realistic goal.

ACKNOWLEDGMENTS

The author's experience on the boll weevil was gained when employed
by Texas A & M University and Centro de Investigaciones Ecologicas del
Sureste at Tapachula, Chiapas, Mexico. Work at Texas A & M University
leading to some of the views expressed in this publication was supported in
part by the National Science Foundation and the Environmental Protection
Agency, through a grant (NSF GB-34718) to the University of California;
this work was in cooperation with the Agricultural Research Service of the
USDA and was funded in part under USDA Cooperative Agreement No.
12-14-100-11, 194(33). Work in Mexico was supported through a grant by
the Organization of American States. Any opinions, findings, conclusions,
and/or recommendations contained in this publication are those of the
author and do not necessarily reflect the views of past or present institutions
of employment or the granting agencies. The survey of the literature pertain-
ing to this review was completed in March 1982. I extend my sincere thanks
to the many colleagues who provided reprints and other material used for
the review. A very special thanks goes to Harry Dewey and Carolyn Wong
for assisting with the literature search; to Carl B. Huffaker, Sheila A. Mulvi-
hill, John H. Perkins, Jacob R. Phillips, and Don R. Rummel for reviewing
the manuscripts; and to Elane Bottrell and Pua Mitchell for typing the
manuscripts.

REFERENCES

Adkisson, P.L., Niles, G.A., Walker, J.K., Bird, L.S. and Scott, H.B., 1982. Controlling
cotton's insect pests: a new system. Science, 216: 19-22.
Adkisson, P.L., Rummel, D.R., Sterling, W.L. and Owen, W.L., Jr., 1966. Diapause boll
weevil control: a comparison of two methods. Texas A & M University, Texas Agri-
cultural Experiment Station, Bulletin 1054, 11 pp.
ARS, 1976. Boll Weevil Suppression, Management, and Elimination Technology. Proceed-
ings of a Conference, 13-15 February 1974, Memphis, TN. United States Department
of Agriculture, Agricultural Research Service, ARS-S-71, 172 pp.
Beckham, C.M., 1957. Hibernating sites of the boll weevil in relation to a small, Georgia
Piedmont cotton field. J. Econ. Entomol., 50: 833-834.
Beckham, C.M. and Morgan, L.W., 1960. On the flight distance of the boll weevil. J.
Econ. Entomol., 53: 681-682.
Benedict, J.H. and George, D.M., 1979. A bibliography of host plant resistance literature
for the boll weevil, Anthonomus grandis. Bull. Entomol. Soc. Am., 25: 19-23.
Bodegas Valera, P.R., Flores Garcfa, R. and de Coss Flores, M.E., 1977. Aspectos de
interes sobre las hospederas alternantes del picudo del algodonera A. grandis y avances
en la investigacion respectiva en el Soconusco, Chiapas, Mexico. Centro de Investiga-
ciones Ecologicas del Sureste, Boletfn de Informacion 3, 14 pp.
Bottrell, D.G., 1976. Biological control agents of the boll weevil. In: Boll Weevil Suppres-
sion, Management, and Elimination Technology. Proceedings of a conference, 13-15
February, 1974, Memphis, TN. United States Department of Agriculture, Agricul-
tural Research Service, ARS-S'71, pp. 22-25.
Bottrell, D.G. and Adkisson, P.L., 1977. Cotton insect pest management. Annu. Rev.
Entomol., 22: 451-481.
Bottrell, D.G. and Moody, D.S., 1973. Factors involved in the late-season emigration of
boll weevil from cotton. In: Principles, Strategies and Tactics of Pest Population
Regulation and Control in the Cotton Ecosystem. Annual Project Report, Texas A &
M University, Texas Agricultural Experiment Station, pp. 57-69.
Bottrell, D.G., White, J.R., Moody, D.S. and Hardee, D.D., 1972. Overwintering habitats
of the boll weevil in the Rolling Plains of Texas. Environ. Entomol., 1: 633-638.
Brazzel, J.R., 1959. The effect of late-season applications of insecticides on diapausing
boll weevils. J. Econ. Entomol., 52: 1042-1045.
Brazzel, J.R. and Newsom, L.D., 1959. Diapause in Anthonomus grandis Boh. J. Econ.
Entomol., 52: 603-611.
Burke, H.R., 1968. Geographic variation and taxonomy of Anthonomus grandis Bohe-
man. Dept. Entomol., Texas A & M University, ERD Contract Number 12-14-100·7733
(33),152 pp.
Burke, H.R. and Cate, J.R., 1979. A new species of Mexican Anthonomus related to the
boll weevil (Coleoptera: Curculionidae). Ann. Entomol. Soc. Am., 72: 189-192.
Burke, H.R. and Clark, W.E., 1976. Cienfuegosia drummondii as a host of the boll weevil,
Anthonomus grandis, in south Texas. In: Boll Weevil Suppression, Management, and
Elimination Technology. Proceedings of a Conference, 13-15 February 1974, Mem-
phis, TN. United States Department of Agriculture, Agricultural Research Service,
ARS-S-n, pp. 12-21.
Cate, J.R. and Clark, W.E., 1978. Report on foreign exploration for parasites of boll
weevil, Anthonomus grandis Boheman, October, 1976. Texas A & M University, Dept.
Entomol. (unpublished report).
Cate, J.R., Curry, G.L. and Feldman, R.M., 1979. A model for boll weevil ovipositional
site selection. Environ. Entomol., 8: 917-921.
Coakley, J.M., Maxwell, F.G. and Jenkins, J.N., 1969. Influence of feeding, oviposition,
and egg and larval development of the boll weevil on abscission of cotton squares. J.
Econ. Entomol., 62: 244-245.
Cook, O.F., 1904a. Evolution of weevil - resistance in cotton. Science, 20: 666-670.
Cook, O.F., 1904b. Report on the habits of the kelep, or Guatemalan cotton-boll-weevil
ant. United States Department of Agriculture, Bureau of Entomology, Bulletin, 49,
15 pp.
Cross, W.H., 1973. Biology, control, and eradication of the boll weevil. Annu. Rev. En-
tomol., 18: 17-46.
Cross, W.H. and Chesnut, T.L., 1971. Arthropod parasites of the boll weevil, Anthono-
mus grandis: I. an annotated list. Ann. Entomol. Soc. Am., 64: 516-527.
Cross, W.H., Lukefahr, M.J., Fryxell, P.A. and Burke, H.R., 1975. Host plants of the boll
weevil. Environ. Entomol., 4: 19-26.
Curry, G.L., Sharpe, P.J.H., DeMichele, D.W. and Cate, J.R., 1980. Towards a manage-
ment model of the cotton-boll weevil ecosystem. J. Environ. Manage., 11: 187-223.
Curry, G.L., Cate, J.R. and Sharpe, P.J.H., 1982. Cotton bud drying: contributions to
boll weevil mortality. Environ. Entomol., 11: 344-350.
Davich, T.B., Hardee, D.D. and Alcala, J.M., 1970. Long-range dispersal of boll weevils
determined with wing traps baited with males. J. Econ. Entomol., 63: 1706,_1708.
De Coss Flores, M.E., Bodegas Valera, P.R. and Flores Garcfa, R., 1977. Parasites locali-
zados para las diferentes especies plagas del algodonero en la Zona del Boconusco,
Chiapas, Mexico. Centro de Investigaciones Ecologicas del Sureste, Boletfn de Infor-
macion 1, 11 pp.
DeMichele, D.W. and Bottrell, D.G., 1976. Systems approach to cotton insect pest
management. In: J.L. Apple and R.F. Smith (Editors), Integrated Pest Management.
Plenum Press, New York, NY, pp. 107-132.
DeMichele, D.W., Curry, G.L., Sharpe, P.J.H. and Barfield, C.S., 1976. Cotton bud
drying: a theoretical model. Environ. Entomol., 5: 1011-1016.
Fenton, F.A., 1952. Field Crop Insects. Macmillan, New York, NY, 405 pp.
Flores Garcfa, R., Bodegas Valera, P.R. and de Coss Flores, M.E., 1977. Determinacion
del estado fisiol6gico en el cual el picudo del algodonero Anthonomus grandis, atra-
viesa la epoca seca en la Zona del Soconusco, Chiapas, Mexico. Centro de Investiga-
ciones Ecol6gicas del Sureste, Boletfn de Informacion, 2,9 pp,
Fye, R.E., 1968a. The thurberia weevil in Arizona. J. Econ. Entomol., 61: 1264-1268.
Fye, R.E., 1968b. Spread of the boll weevil by drainage water and air currents. J. Econ.
Entomol., 61: 1418-1424.
Fye, R.E., 1974. Modeling of cotton insect populations. United States Department of
Agriculture, Agricultural Research Service, ARS W-14, 6 pp.
Fye, R.E. and Leggett, J.E., 1969. Winter survival in Arizona of thurberia weevils released
from thurberia bolls. J. Econ. Entomol., 62: 467-470.
Fye, R.E. and Parencia, C.R., Jr., 1972. The boll weevil complex in Arizona. United
States Department of Agriculture, Agricultural Research Service, Production Research
Report 139, 24 pp.
Fye, R.E., McMillian, W.W., Walker, R.L. and Hopkins, A.R., 1959. The distance into
woods along a cotton field at which the boll weevil hibernates. J. Econ. Entomol., 52:
310-312.
Gaines, J.C., 1932. Studies on the progress of boll weevil infestation at various distances
from hibernation quarters. J. Econ. Entomol., 25: 1181-1187.
Gaines, R.C., 1959. Ecological investigations of the boll weevil, Tallulah, Louisiana,
1915-1958. United States Department of Agriculture, Agricultural Research Service,
Technical Bulletin 1208, 20 pp.
Gilbert, N., Gutierrez, A.P., Frazer, B.D. and Jones, R.E., 1976. Ecological Relationships.
W.H. Freeman, Reading, 157 pp.
Graham, H.M., Hernandez, N.S., Jr., Llanes, J.R. and Tamayo, J.A., 1978. Overwintering
habitats of the boll weevil in the Lower Rio Grande Valley, Texas. Environ. Entomol.,
7: 345-348.
Guerra, A.A., Garcia, R.D. and Tamayo, J.A., 1982. Physiological activity of the boll
weevil during the fall and winter in subtropical areas of the Rio Grande Valley of
Texas. J. Econ. Entomol., 75: 11-15.
Gutierrez, A.P., Falcon, L.A., Loew, W., Leipzig, P.A. and van den Bosch, R., 1975. An
analysis of cotton production in California: a model for Acala cotton and the effects
of defoliators on its yields. Environ. Entomol., 4: 125-136.
Gutierrez, A.P., Wang, Y. and Daxl, R., 1979. The interaction of cotton and cotton boll
weevil (Coleoptera: Curculionidae) - a study of co-adaptation. Can. Entomol., 111:
357-366.
Gutierrez, A.P., DeMichele, D.W., Ying, W., Curry, G.L., Skeith, R. and Brown, L.G.,
1980. The systems approach to research and decision making for cotton pest control.
In: C.B. Huffaker (Editor), New Technology of Pest Control. Wiley, New York, NY,
pp.155-186.
Hardee, D.O., Cross, W.H. and Mitchell, E.B., 1969. Male boll weevils are more attractive
than cotton plants to boll weevils. J. Econ. Entomol., 62: 165-169.
Harned, R.W., 1910. Boll weevil in Mississippi, 1909. Agricultural College, Mississippi
Agricultural Experiment Station, Bulletin 139, 43 pp.
Helms, D., 1980. Revision and reversion: changing cultural control practices for the
cotton boll weevil. Agric. Hist., 54: 108-125.
Hinds, W.E., 1916. Boll weevil in Alabama. Alabama Polytechnic Institute, Alabama
Agricultural Experiment Station, Bulletin 188, 64 pp.
Howard, L.O., 1894. A new cotton insect in Texas. United States Department of Agri-
culture, Division of Entomology, Insect Life, 7: 273.
Howell, A.H., 1907. The relation of birds to the cotton boll weevil. United States Depart-
ment of Agriculture, Biological Survey, Bulletin 29, 31 pp.
Huffaker, C.B., 1972. Ecological management of pest systems: In: J.A. Behnke (Editor),
Challenging Biological Problems: Directions Toward Their Solution. Oxford Univer-
sity Press, New York, NY, pp. 313-342.
Huffaker, C.B., Luck, R.F. and Messenger, P.S., 1977. The ecological basis of biological
control. In: Proc, XV Internatl. Congr. Entomol., 19-27 August 1976, Washington,
DC. Entomological Society of America, College Park, MD, pp. 560-586.
Hunter, W.O., 1909. The boll weevil problem, with special reference to means of reducing
damage. United States Department of Agriculture, Farmers' Bulletin 344, 46 pp.
Hunter, W.D. and Coad, B.R., 1923. The boll-weevil problem. United States Department
of Agriculture, Farmers' Bulletin 1329, 30 pp.
Hunter, W.D. and Hinds, W.E., 1905. The Mexican cotton boll weevil. United States
Department of Agriculture, Bureau of Entomology, Bulletin 51, 181 pp.
Isely , D., 1926. Early summer dispersion of boll weevil with special reference to dusting.
University of Arkansas, Agricultural Experiment Station, Bulletin 204, 17 pp.
Isely, D., 1930. Control of the boll weevil in winter. University of Arkansas, Agricultural
Experiment Station, Bulletin 257: 54-56.
Jenkins, J.N. and Parrott, W.L., 1971. Effectiveness of frego bract as a boll weevil resis-
tance character in cotton. Crop Sci., 11: 739-743.
Johnson, C.G., 1969. Migration and Dispersal of Insects by Flight. Methuen, London,
763 pp.
Johnson, W.L., Cross, W.H., McGovern, W.L. and Mitchell, H.C., 1973. Biology of Hetero-
laccus grandis in a laboratory culture and its potential as an introduced parasite of the
boll weevil in the United States. Environ. Entomol., 2: 112-118.
Johnson, W.L., Cross, W.H. and McGovern, W.L., 1976. Long-range dispersal of marked
boll weevils in Mississippi during 1974. Ann. Entomol. Soc. Am., 69: 421-422.
Jones, J.E., Weaver, J.B., Jr. and Schuster, M.F., 1978. Host plant resistance to the boll
weevil. In: The Boll Weevil: Management Strategies. University of Arkansas, Arkansas
Agricultural Experiment Station, Southern Cooperative Series, Bulletin 228, pp,
50-73.
Jones, J.W., Bowen, H.D., Stinner, R.E., Bradley, J.R., Jr., Sowell, R.S. and Bacheler,
J.S., 1975. Female boll weevil oviposition and feeding processes: a simulation model.
Environ. Entomol., 4: 815-821.
Klages, K.H.W., 1942. Ecological Crop Geography. Macmillan, New York, NY, 615 pp.
Lacewell, R.D., Bottrell, D.G., Billingsley, R.V., Rummel, D.R. and Larson, J.L., 1974.
Impact of the Texas High Plains diapause boll weevil control program. Texas A & M
University, Texas Agricultural Experiment Station, MP-1165, 16 pp.
Lincoln, C. and Waddle, B.A., 1966. Insect resistance of frego-type cotton. Arkansas
Farm Res., 15: 5.
Lloyd, E.P., Tingle, F.C., McCoy, J.R. and Davich, T.B., 1966. The reproduction-diapause
approach to population control of the boll weevil. J. Econ. Entomol., 59: 813-816.
Mally, F.W., 1901. The Mexican cotton-boll weevil. United States Department of Agri-
culture, Farmers' Bulletin 130, 29 pp.
McGovern, W.L. and Cross, W.H., 1976. Affects of two cotton varieties on levels of boll
weevil parasitism (Col.: Curculionidae). Entomophaga, 21: 123-125.
NRC, 1981. Cotton boll weevil: an evaluation of USDA programs. A report prepared by
the Committee on Cotton Insect Management of the National Research Council.
National Academy Press, Washington, DC, 130 pp.
OTA, 1979. Present and future pest management strategies in the control of cotton and
sorghum pests in Texas. In: Pest Management Strategies, Volume II-Working Papers.
Congress of the United States, Office of Technology Assessment, United States
Government Printing Office, Washington, DC, pp. i-vi + 1-75.
Perkins, J.H., 1983. The boll weevil in North America: scientific conflicts over manage-
ment of environmental resources. Agric, Ecosystems Environ., 10: 217-245.
Pfrimmer, T.R. and Merkl, M.E., 1981. Boll weevil: winter survival in surface woods trash
in Mississippi. Environ. Entomol., 10: 419-423.
Phillips, J.R., 1976. Diapause as it relates to the boll weevil Anthonomus grandis Bohe-
man. In: Boll Weevil Suppression, Management, and Elimination Technology. Proceed-
ings of a Conference, 13-15 February 1974, Memphis, TN. United States Department
of Agriculture, Agricultural Research Service, ARS-S-71, pp. 10-11.
Phillips, J.R., Gutierrez, A.P. and Adkisson, P.L., 1980. General accomplishments to-
ward better insect control in cotton. In: C.B. Huffaker (Editor), New Technology of
Pest Control. Wiley, New York, NY, pp. 123-153.
Pierce, W.D., 1912. The insect enemies of the cotton boll weevil. United States Depart-
ment of Agriculture, Bureau of Entomology, Bulletin 100, 99 pp.
Pruitt, G.R., Rummel, D.R., Wade, L.J. and White, J.R., 1978. Effects of a long term
suppression program on boll weevil susceptibility to malathion. Southwest. Entomol.,
3: 215-218.
Ridgway, R.L., Hollingsworth, J.P. and Bull, D.L., 1976. Efficiency of boll weevil phero-
mone traps. In: Detection and Management of the Boll Weevil with Pheromone. Texas
A & M University, Texas Agricultural Experiment Station, Research Monograph 8:
16-19.
Roach, S.H., Ray, L., Taft, H.M. and Hopkins, A.R., 1971. Wing traps baited with male
boll weevils for determining spring emergence of overwintered weevils and subsequent
infestations in cotton. J. Econ. Entomol., 64: 107-110.
Rummel, D.R. and Adkission, P.L., 1970. Distribution of boll weevil-infested cotton
fields in relation to overwintering habitats in the High and Rolling Plains in Texas.
J. Econ. Entomol., 63: 1906-1909.
Rummel, D.R. and Bottrell, D.G., 1976. Seasonally related decline in response of boll
weevils to pheromone traps during mid-season. Environ. Entomol., 5: 783-787.
Rummel, D.R. and Frisbie, R.E., 1978. Suppression of potentially overwintering boll
weevils as a pest management practice. In: The Boll Weevil: Management Strategies.
Arkansas, Agricultural Experiment Station, Southern Cooperative Series, Bulletin
228, pp. 39-49.
Rummel, D.R., Bottrell, D.G., Adkisson, P.L. and McIntyre, R.C., 1975a. An appraisal
of a 10-year effort to prevent the westward spread of the boll weevil. Bull. Entomol.
Soc. Am., 21: 6-11.
Rummel, D.R., White, J.R. and Wade, L.J., 1975b. Late season immigration of boll wee-
vils into an isolated cotton plot. J. Econ. Entomol., 68: 616-618.
Rummel, D.R., Jordan, L.B., White, J.R. and Wade, L.J., 1977. Seasonal variation in the
height of boll weevil flight. Environ. Entomol., 6: 674-678.
Rummel, D.R., White, J.R. and Pruitt, G.R., 1978. A wild feeding host of the boll weevil
in West Texas. Southwest. Entomol., 3: 171-175.
Rummel, D.R., White, J.R., Carroll, S.C. and Pruitt, G.R., 1980. Pheromone trap index
system for predicting need for overwintered boll weevil control. J. Econ. Entomol.,
73: 806-810.
Slosser, J.E. and Boring, E.P., III, 1980. Shelterbelts and boll weevils: a control strategy
based on management of overwintering habitat. Environ. Entomol., 9: 1-6.
Smith, R.F. and van den Bosch, R., 1967. Integrated control. In: W.W. Kilgore and
R.L. Doutt (Editors), Pest Control-Biological, Physical, and Selected Chemical
Methods. Academic Press, New York, NY, pp. 295-340.
Smith, R.F., Apple, J.L. and Bottrell, D.G., 1976. The origins of integrated pest manage-
ment concepts for agricultural crops. In: J.L. Apple and R.F. Smith (Editors), Inte-
grated Pest Management. Plenum Press, New York, NY, pp. 1-16.
Sterling, W.L., 1971. Winter survival of the boll weevil in the High and Rolling Plains
of Texas. J. Econ. Entomol., 64: 39-41.
Sterling, W.L., 1978. Fortuitous biological suppression of the boll weevil by the red
imported fire ant. Environ. Entomol., 7: 564-568.
Sterling, W.L. and Adkisson, P.L., 1971. Seasonal biology of the boll weevil in the High
and Rolling Plains of Texas as compared with previous biological studies of this insect.
Texas A & M University, Texas Agricultural Experiment Station, MP 993, 12 pp.
Sterling, W.L. and Adkisson, P.L., 1978. Population dynamics of the boll weevil inhabit-
ing the High and Rolling Plains of Texas. Environ. Entomol., 7: 439-444.
Stern, V.M., 1973. Economic thresholds. Annu. Rev. Entomol., 18: 259-280.
Taft, H.M. and Jernigan, C.E., 1964. Elevated screens for collecting boll weevils flying
between hibernation sites and cottonfields. J. Econ. Entomol., 57: 773-775.
Townsend, C.H.T., 1895. Report on the Mexican cotton-boll weevil in Texas (Anthono-
mus grandis Boh.). United States Department of Agriculture, Division of Entomology,
Insect Life, 7: 295-309.
Wade, L.J. and Rummel, D.R., 1978. Boll weevil immigration into winter habitat and
subsequent spring and summer emergence. J. Econ. Entomol., 71: 173-178.
Walker, J.K., Jr., 1966. The relationship of the fruiting of the cotton plant and over-
wintered boll weevils to the FJ generation. J. Econ. Entomol., 59: 323-326.
Walker, J.K., Jr. and Bottrell, D.G., 1970. Infestations of boll weevils in isolated plots
of cotton in Texas, 1960-69. J. Econ. Entomol., 63: 1646-1650.
Walker, J.K., Jr. and Niles, G.A., 1971. Population dynamics of the boll weevil and
modified cotton types-implications for pest management. Texas A & M University,
Texas Agricultural Experiment Station, Bulletin 1109, 14 pp.
Warner, R.E., 1966. Taxonomy of the subspecies of Anthonomus grandis (Coleoptera:
Curculionidae). Ann. Entomol. Soc. Am., 59: 1073-1088.
Way, M.J., 1973. Objectives, methods and scope of integrated control. In: P.W. Geier,
L.R. Clark, D.J. Anderson and H.A. Nix (Editors), Insects: Studies in Population
Management. Ecological Society of Australia, Memoirs 1, Canberra, pp. 137-152.
White, J.R. and Rummel, D.R., 1978. Emergence profile of overwintered boll weevils
and entry into cotton. Environ. Entomol., 7: 7-14.

You might also like