Get 20M+ Full-Text Papers For Less Than $1.50/day. Start a 14-Day Trial for You or Your Team.

Learn More →

Boll Weevil Eradication: A Success Story of Science in the Service of Policy and Industry

Boll Weevil Eradication: A Success Story of Science in the Service of Policy and Industry Abstract The boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), is an infamous pest of commercially cultivated cotton, Gossypium hirsutum L. (Malvales: Malvaceae). Once the most important agricultural pest in the United States, the boll weevil spurred an unprecedented mobilization of federal support and cooperation among stakeholders, culminating in the eventual eradication of the species from 98% of its invasive range in the United States. The US Boll Weevil Eradication Program, a joint effort of local, state, and federal governments and agencies, university and agency researchers, and the cotton producers themselves, is a prime example of a successful implementation of a community-wide integrated pest management (IPM). The program also dramatically reduced the use of insecticides in cotton production which led to further positive economic outcomes for producers and reduced the non-target impacts from those chemicals. Though the boll weevil has been mostly eradicated in the United States, the insect remains one of the most important and impactful cotton pests in Central and South America. In this review, we will revisit the agro-economic history of the boll weevil and examine the success of the US Boll Weevil Eradication Program. In doing so, we will learn how we can apply those lessons to boll weevil management abroad and community-wide management of other agricultural or invasive pests. Finally, we will conclude with a brief summary of the ongoing science that continues in service of eradication today. The boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), is an important economic pest of the commercially cultivated varieties of its preferred host, Gossypium hirsutum L. (Malvales: Malvaceae), commonly known as upland cotton. A. grandis is one of the most important pests in the agricultural history of the United States (US), and it remains a persistent problem in cotton-growing regions of Central and South America. In this review, we will revisit the history of the boll weevil in the United States, examine the US Boll Weevil Eradication Program and what made it successful and consider lessons from the boll weevil story that can be applied to area-wide pest management and invasive species mitigation today. Biology and Taxonomy of the Boll Weevil Though both male and female weevils feed on the developing flowers (squares) and fruits (bolls) of the cotton plant, yield loss primarily results from the abscission of squares and bolls that have been oviposited in, which ultimately reduces the amount of lint that the plant produces (Ramalho and Wanderley 1996). Boll weevil larvae complete development inside abscised squares and bolls and emerge to reinfest nearby cotton plants. Total development time from egg to adult ranges from 12 to 90 d depending on the temperature of the surrounding environment, and the adult life span has been reported to be around 40 d (Cross 1973, Ramalho and Wanderley 1996). Under favorable conditions, there may be as many as 5–6 generations in a single year. Some populations of adult weevils may enter a facultative diapause and overwinter in regrowth or volunteer cotton, or even leaf litter (Cross 1973). Consequentially, populations may grow large over the course of a single growing season and remain established in the following year. There are three morphological variants of boll weevil that have been described and historically accepted in North America (Warner 1966, Fye 1968, Cross 1973, Burke et al. 1986). The variant of greatest agro-economic importance in the United States is undoubtedly the southeastern boll weevil, A. g. grandis. Though the subspecies is known primarily as a pest of cotton, it may also utilize other malvaceous plants when G. hirsutum is not available (Cross et al. 1975). A second variant, the Thurberia weevil (A. g. thurberiae), is described as having an extremely restricted geographic range in correlation with the distribution of its preferred host plant, Arizona wild cotton (G. thurberi Tod. [Malvales: Malvaceae]). The Thurberia weevil is typically regarded as a non-pest variant but may incidentally or opportunistically utilize G. hirsutum. The third variant of the species is commonly referred to as the Mexican boll weevil and is found throughout Mexico, Central America, and Cuba. Though A. g. grandis and A. g. thurberiae are widely regarded as distinct subspecies, studies have shown that the morphological characteristics that are used to define the groups are labile to rearing condition and may not be reliable for diagnostic purposes (Warner 1966, Fye 1968, Burke et al. 1986, Roehrdanz 2001, Barr et al. 2013). The Mexican boll weevil has never been given any formal subspecies designation and has generally been considered to be an intermediate form of the two other variants (Cross 1973, Cross et al. 1975, Burke et al. 1986). Some genetic studies have suggested that the subspecies taxonomy is altogether inappropriate for this species and that morphologically distinct groups should be regarded simply as different genetic populations (Kuester et al. 2012, Alvarado et al. 2017, Raszick et al. 2021). Nonetheless, the variant most relevant to this review has been mainly referred to as the southeastern boll weevil (A. g. grandis) throughout the literature, so that nomenclature will be used. A Brief Agro-Economic History of the Boll Weevil in the United States Native to Meso-America, the boll weevil has been infesting cultivated G. hirsutum since at least 900 AD (Warner and Smith 1968). More recently, the expansion of the cultivation of cotton across the Americas has allowed for major range expansions of significant agro-economic impact. The first of these range expansions occurred in the late 1800s toward the north of the species’ native range into North America. This range expansion was likely enabled by increased cotton production in southern Texas and northern Mexico in the 1860s and eventually led to the first record of A. grandis in the United States in 1892 (Howard 1894). Reported from extreme southern Texas, near Brownsville, the boll weevil rapidly spread eastward throughout the US Cotton Belt, reaching the Carolinas and Virginia by 1922 (Fig. 1). In subsequent years, boll weevils also began appearing in cultivated cotton across the southwest and west coast of the United States. It remains unclear if these western infestations were the result of a westward dispersal of the populations infesting the US Cotton Belt, the northward dispersal of boll weevils from cultivation areas in Mexico, the infestation of G. hirsutum by Thurberia weevils, or some combination of the above. Regardless, weevils began infesting cotton in Arizona as early as 1920 (though infestations were mainly intermittent), and California reported an established population in 1982 (Clark 2001, Neal and Antilla 2001). New Mexico was largely spared until the 1990s but did eventually develop infestations in the southern and eastern portions of the state (Pierce et al. 2001). Fig. 1. Open in new tabDownload slide Map illustrating the spread of the boll weevil throughout the US Cotton Belt (1892–1921). Adapted from Coad et al. (1922). Fig. 1. Open in new tabDownload slide Map illustrating the spread of the boll weevil throughout the US Cotton Belt (1892–1921). Adapted from Coad et al. (1922). The introduction of A. g. grandis to Texas in 1892 marked the beginning of a long and costly struggle for the US cotton industry (Fig. 2). Initial yield losses in the late 1800s and early 1900s were estimated to be around 50% nationwide (Lange et al. 2009). From 1921 to 1925, yield losses were estimated at more than 17% (Haney 2001). Despite the fact that the US cotton industry as a whole was expanding during the late 1800s and early 1900s, the establishment of the pest in the US Cotton Belt devastated local economies. In some areas, cotton producers simply gave up and abandoned their land or were forced to adopt alternative crops. Indeed, many attribute the diversification of the economy of the American South to the invasion of the boll weevil (Giesen 2011, Lange et al. 2009). By 1973, boll weevil damage was causing losses of over US$200 million per year, and suppression costs totaled an additional US$75 million annually (Cross 1973). The total economic impact on US cotton has been estimated at more than US$15 billion to as much as US$23 billion after adjusting for inflation (Haney 2001). Further, the spread of the boll weevil threatened whole states’ economies as disputes between states and considerations of quarantines arose. Nonetheless, cotton remained an essential component of the economy in the southern United States. However, the impact of the boll weevil invasion cannot be understated. In some cases, entire towns were displaced as the local economies collapsed, and the early methods of pest suppression came too late for many who had already lost their homes and their way of life (Haney 2001). A slight silver lining of the economic disruption is that educational outcomes for school-aged children in the early years of the invasion may have actually improved as less labor was required as cotton farms shut down (Baker et al. 2020). Fig. 2. Open in new tabDownload slide Timeline of key events in US boll weevil history. Fig. 2. Open in new tabDownload slide Timeline of key events in US boll weevil history. Early Methods of Boll Weevil Control In the years immediately following the initial boll weevil invasion, very little was known about the species’ basic biology, and there were no known control methods (Haney 2001). In the late 1890s, the US Department of Agriculture (USDA) Bureau of Entomology turned to the growers themselves to develop what became known as the “Government Method” for boll weevil management (Haney 2001). Special recognition must be made of the producers in south Texas who were first among US producers to encounter the pest, and USDA scientists worked directly with these front-line growers to identify successful control tactics. The Government Method consisted mainly of cultural control practices intended to minimize the suitable habitat for boll weevil reproduction (Haney 2001). Some key Government Method practices were planting early to avoid the weevil cohort emerging from diapause, the removal, destruction, or plowing under infested or abscised bolls and squares, and the destruction of the cotton crop after harvesting to reduce overwintering habitat (Haney 2001). Some cotton producers were either unwilling or unable to adopt some of the Government Method practices. In particular, the practice of crop destruction after harvest was not widely adopted because additional yield can be obtained from the regrowth of stalks from the previously harvested crop (Haney 2001, Stavinoha and Woodward 2001). Concurrent with the deployment of the Government Method, other public and private entities began developing and testing other potential methods of boll weevil control (Haney 2001). In 1918, the first inorganic insecticide for boll weevil, calcium arsenate, became available. This development marked the beginning of a boll weevil management era defined by the widespread use of broad-spectrum pesticides. Over the next few decades, a variety of inorganic and, later, organic insecticides rotated through the market, but with them came new concerns. One of these concerns was the possible non-target impacts on beneficial insects such as pollinators and the natural enemies of the boll weevil and other cotton pest insects. In the 1920s and 1930s, natural enemies were one of the primary components of biological control upon which farmers relied, and cotton was no different. Biological control was only somewhat effective for the boll weevil (Lincoln 1969, Cross 1973). There are at least 42 known species of arthropods that attack the boll weevil; however, none of these provide complete suppression of local populations at natural levels, though releases of parasitoids have been employed with limited success (Cross and McGovern 1969, Cross et al. 1969b, Chesnut and Cross 1971, Cross and Chesnut 1971). Ironically, one of the best sources of biological control has come from the invasion of another pest species; the red imported fire ant, Solenopsis invicta Buren (Hymenoptera: Formicidae), which has been shown to significantly suppress weevil populations in some areas (Sterling 1978). After the advent of DDT in 1946, cotton production became increasingly dependent on organic pesticides (though DDT itself was not effective for boll weevil control). It has been estimated that during this period (late 1940–1980s), cotton accounted for one-third to one-half of all pesticide use in the United States (Haney 2001). Product and application costs also increased, which created an incentive for the development of alternative methods of control. The possibility of the species developing resistance also incentivized research and development of alternatives. Indeed, tolerance to the organochlorine class of insecticides was observed as early as 1950, and complete resistance was observed later that decade (Rainwater and Gaines 1951, Roussel and Clower 1957, Parencia 1959, Parencia and Cowan 1960). Influence on the USDA and Breakthroughs in Boll Weevil Research The invasion of the boll weevil had a profound effect on the USDA. Though the agency had existed in some form since 1839, the impetus provided by the boll weevil spurred a rapid expansion of entomological research and extension activities by the agency (Haney 2001). In 1902, the first USDA Boll Weevil Laboratory was established, and the agency continued to support growers and researchers throughout the inorganic and organic pesticide eras. Nonetheless, the pest continued to expand its range across the country. In 1958, the National Cotton Council of America (NCC), an organization made up of and representing a broad cross-section of stakeholders in the industry, became unified in its mission and officially recognized the economic impact of the boll weevil (Carter et al. 2001). After this resolution to declare “war” on the boll weevil passed unanimously, the NCC (especially Robert Coker, a producer from South Carolina) went to work and successfully advocated for increased congressional support for boll weevil research (Carter et al. 2001). This led to a major expansion of the USDA’s facilities and capabilities as a whole and directly contributed to the eventual deployment of the US Boll Weevil Eradication Program (Carter et al. 2001, Haney 2001). In the following years, massive strides were made toward understanding the biology of the pest, how to mass rear the insect, and how to control it. One critical breakthrough in boll weevil research was the discovery and description of the species’ facultative diapause and how to exploit it for improved control (Brazzel and Newsom 1959, Brazzel and Hightower 1960, Brazzel et al. 1961). Brazzel and Newsom (1959) found that diapausing boll weevils could be found in the overwintering habitat throughout the calendar year. Brazzel and Hightower (1960) took it a step further and suggested that weevils from subsequent generations could enter diapause while individuals from the previous growing season may have not yet exited diapause. This implied that there could be a multi-generation bank of individuals in the overwintering habitat that could serve as founders for new infestations even after fields had already been treated. The key to combating the weevil was to prevent these source populations from forming. Both studies found that there was a short period of time before and during harvest wherein weevils entered diapause and moved to the overwintering habitat. Brazzel et al. (1961) suggested that a late season pesticide treatment to reduce the overwintering population could dramatically improve outcomes in the following year. This so-called “diapause method” was found to be extremely successful in suppression of boll weevil populations in many areas and allowed for a dramatic reduction in insecticide applications for boll weevil during the growing season (Fye et al. 1968; Rummel and Adkisson 1971; Lloyd et al. 1972, 1966, 1967). A second major breakthrough in boll weevil management was the isolation of the boll weevil aggregation pheromone and the development of specialized boll weevil traps. Early trapping efforts used live male weevils as bait, but the male pheromone eventually became commercially available under the name Grandlure (Plato Industries Inc., Houston, TX). Since then, the use of pheromone-baited traps has been widely adopted as a more practical alternative to using live males (Tumlinson et al. 1969, Hardee et al. 1971). Though not an effective measure of control on their own, baited traps proved to be extremely effective for monitoring populations of boll weevil and for early detection of new infestations, and these traps would become an essential piece of the broader eradication effort (Cross et al. 1969a, Cross 1973). It was also desirable for researchers to have the ability to mass rear boll weevils for experiments. Additionally, the use of sterile insect technique (SIT) was envisioned as a possible solution to the boll weevil problem. The basic idea of SIT is that the mass rearing and subsequent release of sterile males into natural populations will overwhelmingly lead to oviposition of unfertilized eggs, eventually leading to a population crash (McKibben et al. 2001). An SIT approach would have required that millions of boll weevils were reared, sterilized, and released. Before the late 1950s; however, experiments were limited to the use of field-collected weevils, and weevils had proved difficult to rear from small founder populations in a laboratory setting (Betz 1966, Gast and Davich 1966). The mass rearing of boll weevils became feasible when Vanderzant and Davich (1958) published an artificial diet for the species. This development, along with other advances in mechanization, finally allowed for the mass rearing of weevils. Unfortunately, this did not lead to success with SIT. At the time that SIT was being considered as a method of control for boll weevil, sterilization was typically carried out via irradiation, and the doses of radiation required to sterilize boll weevils also led to unacceptably high mortality, and other methods of sterilization also failed (Davich and Lindquist 1962, McKibben et al. 2001). The US Boll Weevil Eradication Program: Inception, Implementation, and Success Finally equipped with the necessary tools for effective boll weevil management, the NCC and the cotton scientific community began to see a path towards the complete eradication of the boll weevil from the United States. In 1969, the NCC established a Special Study Committee on Boll Weevil Eradication headed by Robert Coker (Carter et al. 2001). Coker and the committee, made up of producers, industry leaders, and USDA and university scientists, were charged with developing and implementing a large-scale field trial to evaluate the feasibility of eradication (Carter et al. 2001). However, there remained significant policy and funding roadblocks in the way of the implementation of a national program. Though a national program would be administered by the USDA Animal and Plant Health Inspection Service (APHIS), the actual implementation of the program would be done on a state-by-state basis and executed in regional eradication zones. This design was critical because the method of eradication, especially the diapause method, necessitated a community-wide IPM approach. Additionally, the program was to be funded under a cost sharing model wherein 30% of the costs would be incurred by USDA-APHIS and 70% would be borne by states’ producers (It should be noted that in some states, producer contributions to the program would be fully or partially offset by financial support from the state). So, the program necessitated a literal buy-in from producers in each state and a commitment from local producers to adhere to the standards set forth by the USDA-APHIS. Additionally, policy changes were required to give local, state, and federal officials the authority to enforce the standards of the eradication program at the local, state, and federal levels. Finally, producers in a proposed eradication zone had to vote by two-thirds majority on a referendum to permit expansion of the program into that region. These challenges necessitated massive extension, outreach, and education efforts to secure stakeholder support. Through diligent work throughout the mid-1970s, the NCC ultimately secured the necessary federal support and state policy changes for the first full scale eradication trial, and in 1978, the program was launched in Virginia and North Carolina (Carter et al. 2001, Ridgway and Mussman 2001). The trial was initiated in this region because it was the extreme north and eastern portion of the weevil’s range (Tate 2001). Program managers envisioned that, if successful, the program would then move south and west, slowly reducing the weevil’s range (Tate 2001). During this trial, managers employed the diapause method, SIT, and trapping for population monitoring and early detection of infestations. The trial was immensely successful, and by the third year of the trial, the boll weevil was declared eradicated from Virginia (Tate 2001); the modern eradication era had begun. South Carolina growers joined the US Boll Weevil Eradication Program in 1983, leading to the formation of the Carolinas Program and the Southeastern Boll Weevil Eradication Foundation in 1988, which later grew to include Alabama, Florida, and Georgia eradication programs. Producers approved expansion referenda in those states by 1987, and other states soon followed suit (Brashear and Brumley 2001). California and Arizona initiated eradication in the early 1980s in response to an outbreak of weevils that formed in regrowth cotton (Carter et al. 2001). The relationship between the two western states was initially hostile, and California threatened to sue Arizona if the state did not begin eradication (Neal and Antilla 2001). Shortly thereafter, the states established the Southwestern Boll Weevil Eradication Program with cooperators in northwest Mexico, and eradication was completed in the late 1980s. Texas, the country’s largest and leading cotton-producing state, was the first to encounter the boll weevil and maybe the last to finally rid itself of the pest. Considering the size of the state, it was only practical to carry out separate eradication programs in the different cotton growing parts of the state. The Southern Rolling Plains region initiated eradication in 1994, and other regions began their programs throughout the late 1990s. The Lower Rio Grande Valley (LRGV) zone remains the only active eradication zone in the United States. In all 16 other states (and the other zones in Texas) where eradication was implemented, the boll weevil has been completely eliminated. In areas where the boll weevil is considered to be eradicated, critical monitoring continues using pheromone-baited cone traps. These continued efforts protect the investment in eradication and are essential for ensuring that the weevil does not re-establish in those areas. The continued investment helps to ensure a continued success, and undoubtedly, the US Boll Weevil Eradication Program was a success. It is widely regarded as one of the earliest and most successful implementations of a community-wide integrated pest management (IPM) approach. Despite the overwhelming pressure of the pest, the vast area the pest infested, the large and diverse groups of stakeholders that needed to be involved, and the massive level of federal support required to administer the program, the program’s objective was ultimately completed. The key lesson to take away from the success of this program is the importance of stakeholder engagement. While there was admittedly a strong economic incentive for producers to participate (and arguably, the movement began with the growers themselves), the program required significant investment from producers and sweeping changes to their pest management regimes. This required a system where scientists developed information and interacted with producers, who in return provided feedback. This system manifested as the technical advisory committees (at least one of which, in Texas, still functions today), and regular systematic reviews of the program. These efforts promoted producer confidence in the program and allowed scientists to ensure that their recommendations were appropriate for each community. Education also played a key role in soliciting buy-in, as some producers in areas with low levels of infestation were skeptical of the buy-in costs (Barker et al. 2001). Educating producers about how the eradication program helped the industry as a whole and insulated them against changing conditions eventually won skeptics over. To this day, continued engagement and education remain a critical final piece for eradication. Though the design of the traps has changed over the years, pheromone-baited traps remain one of the most important staples of the US monitoring strategy for A. grandis, especially in Texas where active eradication remains underway (Texas Boll Weevil Eradication Foundation, Inc., Harlingen, TX). It is absolutely essential that this monitoring continues to prevent reinfestation of areas where eradication has been successful. In many of those areas, decades have now passed; some modern growers have never experienced a boll weevil infestation and may question why they continue to pay into a program that combats an absent pest. So, outreach must continue to be made to this new generation of potential skeptics to ensure that the incredible investment made into eradication is protected (Barker et al. 2001). The US Boll Weevil Eradication Program: Lessons for Community-Wide IPM Though the diapause biology of the boll weevil certainly allowed for successful exploitation of specific life history traits in a management context, the structure and approach of the eradication program itself were what enabled its success. The program serves as an effective blueprint for other community-wide IPM, illustrating the importance of stakeholder engagement, robust pest management science, and cooperation among groups of stakeholders. In the years immediately following the initial eradication field trial, key personnel including USDA scientists and administrators, NCC representatives, university scientists and extension agents, and state and local community leaders had to work tirelessly to ensure that all of the constituents were unified in their commitment to eradication (Ridgway and Mussman 2001). This commitment had to transcend state lines and, in the case of the Southwestern Boll Weevil Eradication Program, a national border. This is because the ecology of the insect is independent of such borders, and one of the key tactics of the eradication program, the diapause method, needed to be executed in a biologically meaningful area to be effective. In Central and South America, where boll weevil is still a problem, it may very well be the case that international cooperation will be the key to effective management. Communication between scientists and affected communities was also a critical component that led to the program’s success. In the early years of the boll weevil invasion, scientists learned from the growers to develop the Government Method. However, as science uncovered new tactics for control, researchers had to become educators and develop extension and outreach programs that would help producers change their practices. Finally, any such outreach should be prepared to make an economic case for a change in practice. If that fails, policy changes may be required, so governments must also become involved. Cotton producers were initially hesitant to adopt the practice of crop destruction after harvest due to economic concerns. Policy changes were required to enforce that element of the eradication program, and fines may be levied against growers who fail to comply to this day. For emerging agricultural pests or novel biological invasions, perhaps the most important lesson to take from the success of the eradication program is the effectiveness of thorough monitoring. Many of the decisions that the technical advisory committees made were based on trapping data, and insect trapping remains the first line of defense for preventing reinfestations of eradicated areas today. Science in the Continued Service of Eradication Despite the overall success of the US Boll Weevil Eradication Program, eradication remains an active endeavor in the LRGV of Texas. Eradication in this area is complicated by control discrepancies across the border, insecurity and safety issues that affect management, and the fact that this area may be part of the species’ natural range. Furthermore, volunteer cotton (cotton growing in a feral manner outside of cultivation) is a perennial problem in this region. Volunteer cotton provides a refuge habitat for boll weevil populations during the offseason, so the diapause method is not as effective in the LRGV. This continued regional insecurity has led to at least one recent reinfestation of domestic cotton in an area where eradication had previously been successful. This reinfestation has required large-scale, area-wide spraying of malathion, and cost over US$5 million over 3 yr to treat (Texas Boll Weevil Eradication Foundation, Inc., Harlingen, TX). Though that outbreak was restricted to central Texas, such events are of broader concern at the national level because Texas is the single largest corridor of entry for potential reinfestations of US cotton by populations of boll weevil dispersing or otherwise being moved from the LRGV. This highlights the current need for preventative management along the US–Mexico border. Currently, pheromone-baited traps and strict regulations on cotton producers are employed to prevent reinfestations of areas where the boll weevil has been eradicated. In cases where weevils are detected in traps, malathion is used to suppress emerging populations in surrounding areas. This approach has been mostly successful at maintaining eradication, but there have been some issues. Producers in the LRGV are required by law to declare their cotton each year. This requirement is in place so that the Texas Boll Weevil Eradication Program can adequately place monitoring traps, and failure to comply can result in a fine. Nonetheless, there are occasionally producers who are out of compliance, and it is believed that such a producer may have contributed to the outbreak in central Texas. In response to this issue, the eradication program has desired a way to detect undeclared cotton fields using remote sensing or unmanned aerial vehicles (UAVs), and scientists are currently working to develop those methods (Yang et al. 2017). Additionally, other research groups are working on population genomics of the boll weevil to improve diagnostic capabilities for the USDA. Finally, the boll weevil remains one of the most important agricultural pests in Central and South America, and countless scientists there are working to develop the next generation of controls. It may one day be possible to apply an eradication approach to those areas, and the framework upon which to build already exists with the US Boll Weevil Eradication Program. Conclusion The success of the US Boll Weevil Eradication Program can be attributed to thorough and dedicated basic science research, meaningful engagement with key stakeholders, policy support at the federal, state, and local government levels, and stepwise implementation at biologically appropriate spatial scales. For researchers and managers looking for further specifics, the Cross (1973) paper cited within is a seminal boll weevil paper that discusses the ecological history and biology of the species and how control methods interact with those. Additionally, the author recommends the Cotton Foundation Reference Book “Boll Weevil Eradication in the United States through 1999.” The details contained within that text regarding policy and stakeholder engagement would be invaluable to any researcher or manager considering a control program of any scale. The implementation of boll weevil eradication was a historic success, and the program should continue to serve as a blueprint for future IPM and invasive species management for years to come. Acknowledgments I would like to thank James Brumley of the Southeastern Boll Weevil Eradication Program, Dominic Reisig of North Carolina State University, and Ryan Kurtz of Cotton Incorporated, for providing invaluable reference materials for this endeavor. Special recognition goes to Don Parker of the National Cotton Council for providing a copy of Cotton Foundation Reference Book “Boll Weevil Eradication in the United States through 1999” which served as the single most important reference for this article. I would also like to thank Larry Smith (ret.) and the entire Texas Boll Weevil Eradication Foundation for their support and the opportunities to contribute to and learn from the Technical Advisory Committee meetings. I also would like to recognize my postdoctoral advisor, Gregory Sword, for allowing me to spend valuable work hours on this project. Finally, the organizers of this special issue and corresponding symposium, Lina Bernaola and Jocelyn Holt, are thanked for their incredible patience and the invitation to contribute to this unique collection. This manuscript was greatly improved by the comments provided by the editors and anonymous reviewers who donated their time. This project was financially supported by Cotton Incorporated Core Program Grant 16–352 along with funds from the Charles R. Parencia Endowment to the Department of Entomology at Texas A&M University and USDA-APHIS-PPQ Cooperative Agreement AP19PPQS&T00C062 and USDA-APHIS Plant Protection Act Section 7721 agreement AP18PPQS&T00C197. The USDA is an equal opportunity employer. Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the USDA. References Cited Alvarado , A. , R. W. Jones, C. Pedraza-Lara, O. A. Villanueva and E. Pfeiler. 2017 . Reassessment of the phylogeography and intraspecific relationships of western and eastern populations of the boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), in North America . Biol. J. Linn. Soc . 122 : 29 – 45 . Google Scholar Crossref Search ADS WorldCat Baker , R. B. , J. Blanchette, and K. Eriksson. 2020 . Long-run impacts of agricultural shocks on educational attainment: evidence from the boll weevil . J. Econ. Hist . 80 : 136 – 174 . Google Scholar Crossref Search ADS WorldCat Barker , B. , C. H. Parker, A. G. King, and W. Anderson. 2001 . What’s required to finish the job and protect our investment? pp. 509 – 522 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Barr , N. , R. Ruiz-Arce, O. Obregón, R. De Leon, N. Foster, C. Reuter, T. Boratynski, and D. Vacek. 2013 . Molecular diagnosis of populational variants of Anthonomus grandis (Coleoptera: Curculionidae) in North America . J. Econ. Entomol . 106 : 437 – 449 . Google Scholar Crossref Search ADS PubMed WorldCat Betz , N. L . 1966 . Improved laboratory methods for rearing the boll weevil . J. Econ. Entomol . 59 ( 2 ): 374 – 376 . Google Scholar Crossref Search ADS WorldCat Brashear , A. L. , and J. T. Brumley. 2001 . The evolution of program management, USDA/Grower responsibilities and funding in the Southeastern boll weevil eradication program, pp. 77 – 100 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Brazzel , J. R. , and B. G. Hightower. 1960 . A seasonal study of diapause, reproductive activity and seasonal tolerance to insecticides in the boll weevil . J. Econ. Entomol . 53 : 41 – 46 . Google Scholar Crossref Search ADS WorldCat Brazzel , J. R. , and L. D. Newsom. 1959 . Diapause in Anthonomus grandis Boh . J. Econ. Entomol . 52 : 603 – 611 . Google Scholar Crossref Search ADS WorldCat Brazzel , J. R. , T. B. Davich, and L. D. Harris. 1961 . A new approach to boll weevil control . J. Econ. Entomol . 54 : 723 – 730 . Google Scholar Crossref Search ADS WorldCat Burke , H. R. , W. E. Clark, J. R. Cate, and P. A. Fryxell. 1986 . Origin and dispersal of the boll weevil . Bull. Entomol. Soc. Am . 32 : 228 – 238 . Google Scholar OpenURL Placeholder Text WorldCat Carter , F. L. , T. C. Nelson, A. G. Jordan, and J. R. Smith. 2001 . US cotton declares war on the boll weevil, pp. 7 – 24 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Chesnut , T. L. , and W. H. Cross. 1971 . Arthropod parasites of the boll weevil, Anthonomus grandis: 2. Comparisons of their importance in the United States over a period of thirty-eight years . Ann. Entomol. Soc. Am . 64 : 549 – 557 . Google Scholar Crossref Search ADS WorldCat Clark , D. O . 2001 . Boll weevil establishment and eradication in California, pp. 235 – 244 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Coad , B. R. , E. S. Tucker, W. B. Williams, F. F. Bondy, and R. C. Gaines. 1922 . Dispersion of the boll weevil in 1921 . USDA Dept. Circ . 210 : 1 – 3 . Google Scholar OpenURL Placeholder Text WorldCat Cross , W. H . 1973 . Biology, control, and eradication of the boll weevil . Ann. Rev. Entomol . 18 : 17 – 46 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , and T. L. Chesnut. 1971 . Arthropod parasites of the boll weevil, Anthonomus grandis: 1. An annotated list . Ann. Entomol. Soc. Am . 64 : 516 – 527 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , and W. L. McGovern. 1969 . New parasites, Zatropis perdubius and Megaselia aletiae, of the boll weevil, Anthonomus grandis . Ann. Entomol. Soc. Am . 62 : 674 – 674 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , D. D. Hardee, F. Nichols, H. C. Mitchell, E. B. Mitchell, P. M. Huddleston, and J. H. Tumlinson. 1969a . Attraction of female boll weevils to traps baited with males or extracts of males . J. Econ. Entomol . 62 : 154 – 161 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , W. L. McGovern, and H. C. Mitchell. 1969b . Biology of Bracon kirkpatricki and field releases of the parasite for control of the boll weevil . J. Econ. Entomol . 62 : 448 – 454 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , M. J. Lukefahr, P. A. Fryxell, and H. R. Burke. 1975 . Host plants of the boll weevil . Environ. Entomol . 4 : 19 – 26 . Google Scholar Crossref Search ADS WorldCat Davich , T. B. , and D. A. Lindquist. 1962 . Exploratory studies on gamma radiation for sterilization of the boll weevil . J. Econ. Entomol . 55 : 164 – 167 . Google Scholar Crossref Search ADS WorldCat Fye , R. E . 1968 . The thurberia weevil in Arizona . J. Econ. Entomol . 61 : 1264 – 1268 . Google Scholar Crossref Search ADS WorldCat Fye , R. E. , C. L. Cole, F. C. Tingle, A. Stoner, D. F. Martin, and L. F. Curl. 1968 . A reproductive-diapause control program for the boll weevil in the presidio, Texas-Ojinaga, Chihuahua Area, 1965–67 . J. Econ. Entomol . 61 : 1660 – 1666 . Google Scholar Crossref Search ADS WorldCat Gast , R. T. , and T. B. Davich. 1966 . Boll weevils, pp. 405 – 418 . In Insect colonization and mass production . Academic Press , New York, NY, USA . Google Scholar Crossref Search ADS Google Preview WorldCat COPAC Giesen , J. C . 2011 . “The Herald of Prosperity”: tracing the boll weevil myth in Alabama . Agric. Hist . 85 : 24 – 49 . Google Scholar Crossref Search ADS PubMed WorldCat Haney , P. B . 2001 . The cotton boll weevil in the United States: impact on cotton production and the people of the cotton belt, pp. 7 – 24 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Hardee , D. D. , N. M. Wilson, E. B. Mitchell, and P. M. Huddleston. 1971 . Factors affecting activity of grandlure, the pheromone of the boll weevil, in laboratory bioassays . J. Econ. Entomol . 64 : 1454 – 1456 . Google Scholar Crossref Search ADS WorldCat Howard , L. O . 1894 . A new cotton insect in Texas . Insect Life , Washington . 273 . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Kuester , A. P. , R. J. Jones, T. W. Sappington, K. S. Kim, N. B. Barr, R. L. Roehrdanz, P. Senechal, J. D. Nason. 2012 . Population structure and genetic diversity of the boll weevil (Coleoptera: Curculionidae) on Gossypium in North America . Ann. Entomol. Soc. Am . 105 : 902 – 916 . Google Scholar Crossref Search ADS WorldCat Lange , F. , A. L. Olmstead, and P. W. Rhode. 2009 . The impact of the boll weevil, 1892–1932 . J. Econ. Hist . 69 : 685 – 718 . Google Scholar Crossref Search ADS WorldCat Lincoln , C . 1969 . The effect of agricultural practices on insect habitats in a typical delta community . Tall Timbers Conf Ecol Anim Contr Habitat Manage Proc. Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Lloyd , E. P. , F. C. Tingle, J. R. McCoy, and T. B. Davich. 1966 . The reproduction-diapause approach to population control of the boll weevil . J. Econ. Entomol . 59 : 813 – 816 . Google Scholar Crossref Search ADS WorldCat Lloyd , E. P. , F. C. Tingle, M. E. Merkl, E. C. Burt, D. B. Smith, and T. B. Davich. 1967 . Comparison of three rates of application of ultra-low-volume azinphosmethyl in a reproduction-diapause control program against the boll weevil . J. Econ. Entomol . 60 : 1696 – 1699 . Google Scholar Crossref Search ADS WorldCat Lloyd , E. P. , M. E. Merkl, F. C. Tingle, W. P. Scott, D. D. Hardee, and T. B. Davich. 1972 . Evaluation of male-baited traps for control of boll weevils following a reproduction-diapause program in Monroe County, Mississippi . J. Econ. Entomol . 65 : 552 – 555 . Google Scholar Crossref Search ADS WorldCat McKibben , G. H. , E. J. Villavaso, W. L. McGovern, and B. Grefenstette. 2001 . United States department of agriculture – research support, methods development and program implementation, pp. 101 – 136 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher, Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Neal , C. R. , and L. Antilla. 2001 . Boll weevil establishment and eradication in Arizona and Northwest Mexico, pp. 213 – 224 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Parencia , C. R . 1959 . Comparative yields of cotton in treated and untreated plots in insect-control experiments in central texas, 1939–1958 . J. Econ. Entomol . 52 : 757 – 758 . Google Scholar Crossref Search ADS WorldCat Parencia , C. R. , and C. B. Cowan. 1960 . Increased tolerance of the boll weevil and cotton fleahopper to some chlorinated hydrocarbon insecticides in central texas in 1958 . J. Econ. Entomol . 53 : 52 – 56 . Google Scholar Crossref Search ADS WorldCat Pierce , J. B. , C. A. Sutherland, L. J. Dominguez, A. B. Miller, S. S. Sanderson, and J. A. Calvani. 2001 . Boll weevil establishment and eradication in New Mexico, pp. 361 – 374 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Rainwater , C. F. , and J. C. Gaines. 1951 . Seasonal decline in the effectiveness of certain insecticides against boll weevil . J. Econ. Entomol . 44 : 971 – 974 . Google Scholar Crossref Search ADS WorldCat Ramalho , F. S. , and P. A. Wanderley. 1996 . Ecology and management of the boll weevil in South American cotton . Am. Entomol . 42 : 41 – 47 . Google Scholar Crossref Search ADS WorldCat Raszick , T. J. , C. M. Dickens, L. C. Perkin, A. E. Tessnow, C. P. Suh, R. Ruiz-Arce, T. N. Boratynski, M. R. Falco, J. S. Johnston, and G. A. Sword. 2021 . Population genomics and phylogeography of the boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), in the United States, northern Mexico, and Argentina . Evol. Appl . 14 : 1778 – 1793 . Google Scholar Crossref Search ADS PubMed WorldCat Ridgway , R. L. , and H. C. Mussman. 2001 . Integrating science and stakeholder inputs - the pivotal years, pp. 55 – 76 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Roehrdanz , R. L . 2001 . Genetic differentiation of Southeastern boll weevil and thurberia weevil populations of Anthonomus grandis (Coleoptera: Curculionidae) Using Mitochondrial DNA . Ann. Entomol. Soc. Am . 94 : 928 – 935 . Google Scholar Crossref Search ADS WorldCat Roussel , J. S. , and D. F. Clower. 1957 . Resistance to the chlorinated hydrocarbon insecticides in the boll weevil . J. Econ. Entomol . 50 : 463 – 468 . Google Scholar Crossref Search ADS WorldCat Rummel , D. R. , and P. L. Adkisson. 1971 . A two-phased control program designed for maximum suppression of the boll weevil in the high and rolling plains of texas . J. Econ. Entomol . 64 : 919 – 922 . Google Scholar Crossref Search ADS WorldCat Stavinoha , K. D. , and L. A. Woodward. 2001 . Texas boll weevil history, pp. 451 – 502 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Sterling , W. L . 1978 . Fortuitous biological suppression of the boll weevil by the red imported fire ant . Environ. Entomol . 7 : 564 – 568 . Google Scholar Crossref Search ADS WorldCat Tate , J. R . 2001 . Cotton production and the boll weevil in Virginia, pp. 503 – 508 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Tumlinson , J. H. , D. D. Hardee, R. C. Gueldner, A. C. Thompson, P. A. Hedin, and J. P. Minyard. 1969 . Sex pheromones produced by male boll weevil: isolation, identification, and synthesis . Science . 166 : 1010 – 1012 . Google Scholar Crossref Search ADS PubMed WorldCat Vanderzant , E. S. , and T. B. Davich. 1958 . Laboratory rearing of the boll weevil: a satisfactory larval diet and oviposition studies . J. Econ. Entomol . 51 ( 3 ): 288 – 291 . Google Scholar Crossref Search ADS WorldCat Warner , R. E . 1966 . Taxonomy of the Subspecies of Anthonomus grandis (Coleoptera: Curculionidae) . Ann. Entomol. Soc. Am . 59 : 1073 – 1088 . Google Scholar Crossref Search ADS WorldCat Warner , R. E. , and C. E. Smith , Jr. 1968 . Boll weevil found in pre-columbian cotton from Mexico . Science . 162 : 911 – 912 . Google Scholar Crossref Search ADS PubMed WorldCat Yang , C. , C. P. C. Suh, and J. K. Westbrook. 2017 . Early identification of cotton fields using mosaicked aerial multispectral imagery . J. Appl. Remote Sens . 11 : 016008 . Google Scholar Crossref Search ADS WorldCat © The Author(s) 2021. Published by Oxford University Press on behalf of Entomological Society of America. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com. This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model) http://www.deepdyve.com/assets/images/DeepDyve-Logo-lg.png Annals of the Entomological Society of America Oxford University Press

Boll Weevil Eradication: A Success Story of Science in the Service of Policy and Industry

Annals of the Entomological Society of America , Volume Advance Article – Oct 5, 2021

Loading next page...
 
/lp/oxford-university-press/boll-weevil-eradication-a-success-story-of-science-in-the-service-of-5CPqpDNhvR

References (58)

Publisher
Oxford University Press
Copyright
Copyright © 2021 Entomological Society of America
ISSN
0013-8746
eISSN
1938-2901
DOI
10.1093/aesa/saab031
Publisher site
See Article on Publisher Site

Abstract

Abstract The boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), is an infamous pest of commercially cultivated cotton, Gossypium hirsutum L. (Malvales: Malvaceae). Once the most important agricultural pest in the United States, the boll weevil spurred an unprecedented mobilization of federal support and cooperation among stakeholders, culminating in the eventual eradication of the species from 98% of its invasive range in the United States. The US Boll Weevil Eradication Program, a joint effort of local, state, and federal governments and agencies, university and agency researchers, and the cotton producers themselves, is a prime example of a successful implementation of a community-wide integrated pest management (IPM). The program also dramatically reduced the use of insecticides in cotton production which led to further positive economic outcomes for producers and reduced the non-target impacts from those chemicals. Though the boll weevil has been mostly eradicated in the United States, the insect remains one of the most important and impactful cotton pests in Central and South America. In this review, we will revisit the agro-economic history of the boll weevil and examine the success of the US Boll Weevil Eradication Program. In doing so, we will learn how we can apply those lessons to boll weevil management abroad and community-wide management of other agricultural or invasive pests. Finally, we will conclude with a brief summary of the ongoing science that continues in service of eradication today. The boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), is an important economic pest of the commercially cultivated varieties of its preferred host, Gossypium hirsutum L. (Malvales: Malvaceae), commonly known as upland cotton. A. grandis is one of the most important pests in the agricultural history of the United States (US), and it remains a persistent problem in cotton-growing regions of Central and South America. In this review, we will revisit the history of the boll weevil in the United States, examine the US Boll Weevil Eradication Program and what made it successful and consider lessons from the boll weevil story that can be applied to area-wide pest management and invasive species mitigation today. Biology and Taxonomy of the Boll Weevil Though both male and female weevils feed on the developing flowers (squares) and fruits (bolls) of the cotton plant, yield loss primarily results from the abscission of squares and bolls that have been oviposited in, which ultimately reduces the amount of lint that the plant produces (Ramalho and Wanderley 1996). Boll weevil larvae complete development inside abscised squares and bolls and emerge to reinfest nearby cotton plants. Total development time from egg to adult ranges from 12 to 90 d depending on the temperature of the surrounding environment, and the adult life span has been reported to be around 40 d (Cross 1973, Ramalho and Wanderley 1996). Under favorable conditions, there may be as many as 5–6 generations in a single year. Some populations of adult weevils may enter a facultative diapause and overwinter in regrowth or volunteer cotton, or even leaf litter (Cross 1973). Consequentially, populations may grow large over the course of a single growing season and remain established in the following year. There are three morphological variants of boll weevil that have been described and historically accepted in North America (Warner 1966, Fye 1968, Cross 1973, Burke et al. 1986). The variant of greatest agro-economic importance in the United States is undoubtedly the southeastern boll weevil, A. g. grandis. Though the subspecies is known primarily as a pest of cotton, it may also utilize other malvaceous plants when G. hirsutum is not available (Cross et al. 1975). A second variant, the Thurberia weevil (A. g. thurberiae), is described as having an extremely restricted geographic range in correlation with the distribution of its preferred host plant, Arizona wild cotton (G. thurberi Tod. [Malvales: Malvaceae]). The Thurberia weevil is typically regarded as a non-pest variant but may incidentally or opportunistically utilize G. hirsutum. The third variant of the species is commonly referred to as the Mexican boll weevil and is found throughout Mexico, Central America, and Cuba. Though A. g. grandis and A. g. thurberiae are widely regarded as distinct subspecies, studies have shown that the morphological characteristics that are used to define the groups are labile to rearing condition and may not be reliable for diagnostic purposes (Warner 1966, Fye 1968, Burke et al. 1986, Roehrdanz 2001, Barr et al. 2013). The Mexican boll weevil has never been given any formal subspecies designation and has generally been considered to be an intermediate form of the two other variants (Cross 1973, Cross et al. 1975, Burke et al. 1986). Some genetic studies have suggested that the subspecies taxonomy is altogether inappropriate for this species and that morphologically distinct groups should be regarded simply as different genetic populations (Kuester et al. 2012, Alvarado et al. 2017, Raszick et al. 2021). Nonetheless, the variant most relevant to this review has been mainly referred to as the southeastern boll weevil (A. g. grandis) throughout the literature, so that nomenclature will be used. A Brief Agro-Economic History of the Boll Weevil in the United States Native to Meso-America, the boll weevil has been infesting cultivated G. hirsutum since at least 900 AD (Warner and Smith 1968). More recently, the expansion of the cultivation of cotton across the Americas has allowed for major range expansions of significant agro-economic impact. The first of these range expansions occurred in the late 1800s toward the north of the species’ native range into North America. This range expansion was likely enabled by increased cotton production in southern Texas and northern Mexico in the 1860s and eventually led to the first record of A. grandis in the United States in 1892 (Howard 1894). Reported from extreme southern Texas, near Brownsville, the boll weevil rapidly spread eastward throughout the US Cotton Belt, reaching the Carolinas and Virginia by 1922 (Fig. 1). In subsequent years, boll weevils also began appearing in cultivated cotton across the southwest and west coast of the United States. It remains unclear if these western infestations were the result of a westward dispersal of the populations infesting the US Cotton Belt, the northward dispersal of boll weevils from cultivation areas in Mexico, the infestation of G. hirsutum by Thurberia weevils, or some combination of the above. Regardless, weevils began infesting cotton in Arizona as early as 1920 (though infestations were mainly intermittent), and California reported an established population in 1982 (Clark 2001, Neal and Antilla 2001). New Mexico was largely spared until the 1990s but did eventually develop infestations in the southern and eastern portions of the state (Pierce et al. 2001). Fig. 1. Open in new tabDownload slide Map illustrating the spread of the boll weevil throughout the US Cotton Belt (1892–1921). Adapted from Coad et al. (1922). Fig. 1. Open in new tabDownload slide Map illustrating the spread of the boll weevil throughout the US Cotton Belt (1892–1921). Adapted from Coad et al. (1922). The introduction of A. g. grandis to Texas in 1892 marked the beginning of a long and costly struggle for the US cotton industry (Fig. 2). Initial yield losses in the late 1800s and early 1900s were estimated to be around 50% nationwide (Lange et al. 2009). From 1921 to 1925, yield losses were estimated at more than 17% (Haney 2001). Despite the fact that the US cotton industry as a whole was expanding during the late 1800s and early 1900s, the establishment of the pest in the US Cotton Belt devastated local economies. In some areas, cotton producers simply gave up and abandoned their land or were forced to adopt alternative crops. Indeed, many attribute the diversification of the economy of the American South to the invasion of the boll weevil (Giesen 2011, Lange et al. 2009). By 1973, boll weevil damage was causing losses of over US$200 million per year, and suppression costs totaled an additional US$75 million annually (Cross 1973). The total economic impact on US cotton has been estimated at more than US$15 billion to as much as US$23 billion after adjusting for inflation (Haney 2001). Further, the spread of the boll weevil threatened whole states’ economies as disputes between states and considerations of quarantines arose. Nonetheless, cotton remained an essential component of the economy in the southern United States. However, the impact of the boll weevil invasion cannot be understated. In some cases, entire towns were displaced as the local economies collapsed, and the early methods of pest suppression came too late for many who had already lost their homes and their way of life (Haney 2001). A slight silver lining of the economic disruption is that educational outcomes for school-aged children in the early years of the invasion may have actually improved as less labor was required as cotton farms shut down (Baker et al. 2020). Fig. 2. Open in new tabDownload slide Timeline of key events in US boll weevil history. Fig. 2. Open in new tabDownload slide Timeline of key events in US boll weevil history. Early Methods of Boll Weevil Control In the years immediately following the initial boll weevil invasion, very little was known about the species’ basic biology, and there were no known control methods (Haney 2001). In the late 1890s, the US Department of Agriculture (USDA) Bureau of Entomology turned to the growers themselves to develop what became known as the “Government Method” for boll weevil management (Haney 2001). Special recognition must be made of the producers in south Texas who were first among US producers to encounter the pest, and USDA scientists worked directly with these front-line growers to identify successful control tactics. The Government Method consisted mainly of cultural control practices intended to minimize the suitable habitat for boll weevil reproduction (Haney 2001). Some key Government Method practices were planting early to avoid the weevil cohort emerging from diapause, the removal, destruction, or plowing under infested or abscised bolls and squares, and the destruction of the cotton crop after harvesting to reduce overwintering habitat (Haney 2001). Some cotton producers were either unwilling or unable to adopt some of the Government Method practices. In particular, the practice of crop destruction after harvest was not widely adopted because additional yield can be obtained from the regrowth of stalks from the previously harvested crop (Haney 2001, Stavinoha and Woodward 2001). Concurrent with the deployment of the Government Method, other public and private entities began developing and testing other potential methods of boll weevil control (Haney 2001). In 1918, the first inorganic insecticide for boll weevil, calcium arsenate, became available. This development marked the beginning of a boll weevil management era defined by the widespread use of broad-spectrum pesticides. Over the next few decades, a variety of inorganic and, later, organic insecticides rotated through the market, but with them came new concerns. One of these concerns was the possible non-target impacts on beneficial insects such as pollinators and the natural enemies of the boll weevil and other cotton pest insects. In the 1920s and 1930s, natural enemies were one of the primary components of biological control upon which farmers relied, and cotton was no different. Biological control was only somewhat effective for the boll weevil (Lincoln 1969, Cross 1973). There are at least 42 known species of arthropods that attack the boll weevil; however, none of these provide complete suppression of local populations at natural levels, though releases of parasitoids have been employed with limited success (Cross and McGovern 1969, Cross et al. 1969b, Chesnut and Cross 1971, Cross and Chesnut 1971). Ironically, one of the best sources of biological control has come from the invasion of another pest species; the red imported fire ant, Solenopsis invicta Buren (Hymenoptera: Formicidae), which has been shown to significantly suppress weevil populations in some areas (Sterling 1978). After the advent of DDT in 1946, cotton production became increasingly dependent on organic pesticides (though DDT itself was not effective for boll weevil control). It has been estimated that during this period (late 1940–1980s), cotton accounted for one-third to one-half of all pesticide use in the United States (Haney 2001). Product and application costs also increased, which created an incentive for the development of alternative methods of control. The possibility of the species developing resistance also incentivized research and development of alternatives. Indeed, tolerance to the organochlorine class of insecticides was observed as early as 1950, and complete resistance was observed later that decade (Rainwater and Gaines 1951, Roussel and Clower 1957, Parencia 1959, Parencia and Cowan 1960). Influence on the USDA and Breakthroughs in Boll Weevil Research The invasion of the boll weevil had a profound effect on the USDA. Though the agency had existed in some form since 1839, the impetus provided by the boll weevil spurred a rapid expansion of entomological research and extension activities by the agency (Haney 2001). In 1902, the first USDA Boll Weevil Laboratory was established, and the agency continued to support growers and researchers throughout the inorganic and organic pesticide eras. Nonetheless, the pest continued to expand its range across the country. In 1958, the National Cotton Council of America (NCC), an organization made up of and representing a broad cross-section of stakeholders in the industry, became unified in its mission and officially recognized the economic impact of the boll weevil (Carter et al. 2001). After this resolution to declare “war” on the boll weevil passed unanimously, the NCC (especially Robert Coker, a producer from South Carolina) went to work and successfully advocated for increased congressional support for boll weevil research (Carter et al. 2001). This led to a major expansion of the USDA’s facilities and capabilities as a whole and directly contributed to the eventual deployment of the US Boll Weevil Eradication Program (Carter et al. 2001, Haney 2001). In the following years, massive strides were made toward understanding the biology of the pest, how to mass rear the insect, and how to control it. One critical breakthrough in boll weevil research was the discovery and description of the species’ facultative diapause and how to exploit it for improved control (Brazzel and Newsom 1959, Brazzel and Hightower 1960, Brazzel et al. 1961). Brazzel and Newsom (1959) found that diapausing boll weevils could be found in the overwintering habitat throughout the calendar year. Brazzel and Hightower (1960) took it a step further and suggested that weevils from subsequent generations could enter diapause while individuals from the previous growing season may have not yet exited diapause. This implied that there could be a multi-generation bank of individuals in the overwintering habitat that could serve as founders for new infestations even after fields had already been treated. The key to combating the weevil was to prevent these source populations from forming. Both studies found that there was a short period of time before and during harvest wherein weevils entered diapause and moved to the overwintering habitat. Brazzel et al. (1961) suggested that a late season pesticide treatment to reduce the overwintering population could dramatically improve outcomes in the following year. This so-called “diapause method” was found to be extremely successful in suppression of boll weevil populations in many areas and allowed for a dramatic reduction in insecticide applications for boll weevil during the growing season (Fye et al. 1968; Rummel and Adkisson 1971; Lloyd et al. 1972, 1966, 1967). A second major breakthrough in boll weevil management was the isolation of the boll weevil aggregation pheromone and the development of specialized boll weevil traps. Early trapping efforts used live male weevils as bait, but the male pheromone eventually became commercially available under the name Grandlure (Plato Industries Inc., Houston, TX). Since then, the use of pheromone-baited traps has been widely adopted as a more practical alternative to using live males (Tumlinson et al. 1969, Hardee et al. 1971). Though not an effective measure of control on their own, baited traps proved to be extremely effective for monitoring populations of boll weevil and for early detection of new infestations, and these traps would become an essential piece of the broader eradication effort (Cross et al. 1969a, Cross 1973). It was also desirable for researchers to have the ability to mass rear boll weevils for experiments. Additionally, the use of sterile insect technique (SIT) was envisioned as a possible solution to the boll weevil problem. The basic idea of SIT is that the mass rearing and subsequent release of sterile males into natural populations will overwhelmingly lead to oviposition of unfertilized eggs, eventually leading to a population crash (McKibben et al. 2001). An SIT approach would have required that millions of boll weevils were reared, sterilized, and released. Before the late 1950s; however, experiments were limited to the use of field-collected weevils, and weevils had proved difficult to rear from small founder populations in a laboratory setting (Betz 1966, Gast and Davich 1966). The mass rearing of boll weevils became feasible when Vanderzant and Davich (1958) published an artificial diet for the species. This development, along with other advances in mechanization, finally allowed for the mass rearing of weevils. Unfortunately, this did not lead to success with SIT. At the time that SIT was being considered as a method of control for boll weevil, sterilization was typically carried out via irradiation, and the doses of radiation required to sterilize boll weevils also led to unacceptably high mortality, and other methods of sterilization also failed (Davich and Lindquist 1962, McKibben et al. 2001). The US Boll Weevil Eradication Program: Inception, Implementation, and Success Finally equipped with the necessary tools for effective boll weevil management, the NCC and the cotton scientific community began to see a path towards the complete eradication of the boll weevil from the United States. In 1969, the NCC established a Special Study Committee on Boll Weevil Eradication headed by Robert Coker (Carter et al. 2001). Coker and the committee, made up of producers, industry leaders, and USDA and university scientists, were charged with developing and implementing a large-scale field trial to evaluate the feasibility of eradication (Carter et al. 2001). However, there remained significant policy and funding roadblocks in the way of the implementation of a national program. Though a national program would be administered by the USDA Animal and Plant Health Inspection Service (APHIS), the actual implementation of the program would be done on a state-by-state basis and executed in regional eradication zones. This design was critical because the method of eradication, especially the diapause method, necessitated a community-wide IPM approach. Additionally, the program was to be funded under a cost sharing model wherein 30% of the costs would be incurred by USDA-APHIS and 70% would be borne by states’ producers (It should be noted that in some states, producer contributions to the program would be fully or partially offset by financial support from the state). So, the program necessitated a literal buy-in from producers in each state and a commitment from local producers to adhere to the standards set forth by the USDA-APHIS. Additionally, policy changes were required to give local, state, and federal officials the authority to enforce the standards of the eradication program at the local, state, and federal levels. Finally, producers in a proposed eradication zone had to vote by two-thirds majority on a referendum to permit expansion of the program into that region. These challenges necessitated massive extension, outreach, and education efforts to secure stakeholder support. Through diligent work throughout the mid-1970s, the NCC ultimately secured the necessary federal support and state policy changes for the first full scale eradication trial, and in 1978, the program was launched in Virginia and North Carolina (Carter et al. 2001, Ridgway and Mussman 2001). The trial was initiated in this region because it was the extreme north and eastern portion of the weevil’s range (Tate 2001). Program managers envisioned that, if successful, the program would then move south and west, slowly reducing the weevil’s range (Tate 2001). During this trial, managers employed the diapause method, SIT, and trapping for population monitoring and early detection of infestations. The trial was immensely successful, and by the third year of the trial, the boll weevil was declared eradicated from Virginia (Tate 2001); the modern eradication era had begun. South Carolina growers joined the US Boll Weevil Eradication Program in 1983, leading to the formation of the Carolinas Program and the Southeastern Boll Weevil Eradication Foundation in 1988, which later grew to include Alabama, Florida, and Georgia eradication programs. Producers approved expansion referenda in those states by 1987, and other states soon followed suit (Brashear and Brumley 2001). California and Arizona initiated eradication in the early 1980s in response to an outbreak of weevils that formed in regrowth cotton (Carter et al. 2001). The relationship between the two western states was initially hostile, and California threatened to sue Arizona if the state did not begin eradication (Neal and Antilla 2001). Shortly thereafter, the states established the Southwestern Boll Weevil Eradication Program with cooperators in northwest Mexico, and eradication was completed in the late 1980s. Texas, the country’s largest and leading cotton-producing state, was the first to encounter the boll weevil and maybe the last to finally rid itself of the pest. Considering the size of the state, it was only practical to carry out separate eradication programs in the different cotton growing parts of the state. The Southern Rolling Plains region initiated eradication in 1994, and other regions began their programs throughout the late 1990s. The Lower Rio Grande Valley (LRGV) zone remains the only active eradication zone in the United States. In all 16 other states (and the other zones in Texas) where eradication was implemented, the boll weevil has been completely eliminated. In areas where the boll weevil is considered to be eradicated, critical monitoring continues using pheromone-baited cone traps. These continued efforts protect the investment in eradication and are essential for ensuring that the weevil does not re-establish in those areas. The continued investment helps to ensure a continued success, and undoubtedly, the US Boll Weevil Eradication Program was a success. It is widely regarded as one of the earliest and most successful implementations of a community-wide integrated pest management (IPM) approach. Despite the overwhelming pressure of the pest, the vast area the pest infested, the large and diverse groups of stakeholders that needed to be involved, and the massive level of federal support required to administer the program, the program’s objective was ultimately completed. The key lesson to take away from the success of this program is the importance of stakeholder engagement. While there was admittedly a strong economic incentive for producers to participate (and arguably, the movement began with the growers themselves), the program required significant investment from producers and sweeping changes to their pest management regimes. This required a system where scientists developed information and interacted with producers, who in return provided feedback. This system manifested as the technical advisory committees (at least one of which, in Texas, still functions today), and regular systematic reviews of the program. These efforts promoted producer confidence in the program and allowed scientists to ensure that their recommendations were appropriate for each community. Education also played a key role in soliciting buy-in, as some producers in areas with low levels of infestation were skeptical of the buy-in costs (Barker et al. 2001). Educating producers about how the eradication program helped the industry as a whole and insulated them against changing conditions eventually won skeptics over. To this day, continued engagement and education remain a critical final piece for eradication. Though the design of the traps has changed over the years, pheromone-baited traps remain one of the most important staples of the US monitoring strategy for A. grandis, especially in Texas where active eradication remains underway (Texas Boll Weevil Eradication Foundation, Inc., Harlingen, TX). It is absolutely essential that this monitoring continues to prevent reinfestation of areas where eradication has been successful. In many of those areas, decades have now passed; some modern growers have never experienced a boll weevil infestation and may question why they continue to pay into a program that combats an absent pest. So, outreach must continue to be made to this new generation of potential skeptics to ensure that the incredible investment made into eradication is protected (Barker et al. 2001). The US Boll Weevil Eradication Program: Lessons for Community-Wide IPM Though the diapause biology of the boll weevil certainly allowed for successful exploitation of specific life history traits in a management context, the structure and approach of the eradication program itself were what enabled its success. The program serves as an effective blueprint for other community-wide IPM, illustrating the importance of stakeholder engagement, robust pest management science, and cooperation among groups of stakeholders. In the years immediately following the initial eradication field trial, key personnel including USDA scientists and administrators, NCC representatives, university scientists and extension agents, and state and local community leaders had to work tirelessly to ensure that all of the constituents were unified in their commitment to eradication (Ridgway and Mussman 2001). This commitment had to transcend state lines and, in the case of the Southwestern Boll Weevil Eradication Program, a national border. This is because the ecology of the insect is independent of such borders, and one of the key tactics of the eradication program, the diapause method, needed to be executed in a biologically meaningful area to be effective. In Central and South America, where boll weevil is still a problem, it may very well be the case that international cooperation will be the key to effective management. Communication between scientists and affected communities was also a critical component that led to the program’s success. In the early years of the boll weevil invasion, scientists learned from the growers to develop the Government Method. However, as science uncovered new tactics for control, researchers had to become educators and develop extension and outreach programs that would help producers change their practices. Finally, any such outreach should be prepared to make an economic case for a change in practice. If that fails, policy changes may be required, so governments must also become involved. Cotton producers were initially hesitant to adopt the practice of crop destruction after harvest due to economic concerns. Policy changes were required to enforce that element of the eradication program, and fines may be levied against growers who fail to comply to this day. For emerging agricultural pests or novel biological invasions, perhaps the most important lesson to take from the success of the eradication program is the effectiveness of thorough monitoring. Many of the decisions that the technical advisory committees made were based on trapping data, and insect trapping remains the first line of defense for preventing reinfestations of eradicated areas today. Science in the Continued Service of Eradication Despite the overall success of the US Boll Weevil Eradication Program, eradication remains an active endeavor in the LRGV of Texas. Eradication in this area is complicated by control discrepancies across the border, insecurity and safety issues that affect management, and the fact that this area may be part of the species’ natural range. Furthermore, volunteer cotton (cotton growing in a feral manner outside of cultivation) is a perennial problem in this region. Volunteer cotton provides a refuge habitat for boll weevil populations during the offseason, so the diapause method is not as effective in the LRGV. This continued regional insecurity has led to at least one recent reinfestation of domestic cotton in an area where eradication had previously been successful. This reinfestation has required large-scale, area-wide spraying of malathion, and cost over US$5 million over 3 yr to treat (Texas Boll Weevil Eradication Foundation, Inc., Harlingen, TX). Though that outbreak was restricted to central Texas, such events are of broader concern at the national level because Texas is the single largest corridor of entry for potential reinfestations of US cotton by populations of boll weevil dispersing or otherwise being moved from the LRGV. This highlights the current need for preventative management along the US–Mexico border. Currently, pheromone-baited traps and strict regulations on cotton producers are employed to prevent reinfestations of areas where the boll weevil has been eradicated. In cases where weevils are detected in traps, malathion is used to suppress emerging populations in surrounding areas. This approach has been mostly successful at maintaining eradication, but there have been some issues. Producers in the LRGV are required by law to declare their cotton each year. This requirement is in place so that the Texas Boll Weevil Eradication Program can adequately place monitoring traps, and failure to comply can result in a fine. Nonetheless, there are occasionally producers who are out of compliance, and it is believed that such a producer may have contributed to the outbreak in central Texas. In response to this issue, the eradication program has desired a way to detect undeclared cotton fields using remote sensing or unmanned aerial vehicles (UAVs), and scientists are currently working to develop those methods (Yang et al. 2017). Additionally, other research groups are working on population genomics of the boll weevil to improve diagnostic capabilities for the USDA. Finally, the boll weevil remains one of the most important agricultural pests in Central and South America, and countless scientists there are working to develop the next generation of controls. It may one day be possible to apply an eradication approach to those areas, and the framework upon which to build already exists with the US Boll Weevil Eradication Program. Conclusion The success of the US Boll Weevil Eradication Program can be attributed to thorough and dedicated basic science research, meaningful engagement with key stakeholders, policy support at the federal, state, and local government levels, and stepwise implementation at biologically appropriate spatial scales. For researchers and managers looking for further specifics, the Cross (1973) paper cited within is a seminal boll weevil paper that discusses the ecological history and biology of the species and how control methods interact with those. Additionally, the author recommends the Cotton Foundation Reference Book “Boll Weevil Eradication in the United States through 1999.” The details contained within that text regarding policy and stakeholder engagement would be invaluable to any researcher or manager considering a control program of any scale. The implementation of boll weevil eradication was a historic success, and the program should continue to serve as a blueprint for future IPM and invasive species management for years to come. Acknowledgments I would like to thank James Brumley of the Southeastern Boll Weevil Eradication Program, Dominic Reisig of North Carolina State University, and Ryan Kurtz of Cotton Incorporated, for providing invaluable reference materials for this endeavor. Special recognition goes to Don Parker of the National Cotton Council for providing a copy of Cotton Foundation Reference Book “Boll Weevil Eradication in the United States through 1999” which served as the single most important reference for this article. I would also like to thank Larry Smith (ret.) and the entire Texas Boll Weevil Eradication Foundation for their support and the opportunities to contribute to and learn from the Technical Advisory Committee meetings. I also would like to recognize my postdoctoral advisor, Gregory Sword, for allowing me to spend valuable work hours on this project. Finally, the organizers of this special issue and corresponding symposium, Lina Bernaola and Jocelyn Holt, are thanked for their incredible patience and the invitation to contribute to this unique collection. This manuscript was greatly improved by the comments provided by the editors and anonymous reviewers who donated their time. This project was financially supported by Cotton Incorporated Core Program Grant 16–352 along with funds from the Charles R. Parencia Endowment to the Department of Entomology at Texas A&M University and USDA-APHIS-PPQ Cooperative Agreement AP19PPQS&T00C062 and USDA-APHIS Plant Protection Act Section 7721 agreement AP18PPQS&T00C197. The USDA is an equal opportunity employer. Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the USDA. References Cited Alvarado , A. , R. W. Jones, C. Pedraza-Lara, O. A. Villanueva and E. Pfeiler. 2017 . Reassessment of the phylogeography and intraspecific relationships of western and eastern populations of the boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), in North America . Biol. J. Linn. Soc . 122 : 29 – 45 . Google Scholar Crossref Search ADS WorldCat Baker , R. B. , J. Blanchette, and K. Eriksson. 2020 . Long-run impacts of agricultural shocks on educational attainment: evidence from the boll weevil . J. Econ. Hist . 80 : 136 – 174 . Google Scholar Crossref Search ADS WorldCat Barker , B. , C. H. Parker, A. G. King, and W. Anderson. 2001 . What’s required to finish the job and protect our investment? pp. 509 – 522 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Barr , N. , R. Ruiz-Arce, O. Obregón, R. De Leon, N. Foster, C. Reuter, T. Boratynski, and D. Vacek. 2013 . Molecular diagnosis of populational variants of Anthonomus grandis (Coleoptera: Curculionidae) in North America . J. Econ. Entomol . 106 : 437 – 449 . Google Scholar Crossref Search ADS PubMed WorldCat Betz , N. L . 1966 . Improved laboratory methods for rearing the boll weevil . J. Econ. Entomol . 59 ( 2 ): 374 – 376 . Google Scholar Crossref Search ADS WorldCat Brashear , A. L. , and J. T. Brumley. 2001 . The evolution of program management, USDA/Grower responsibilities and funding in the Southeastern boll weevil eradication program, pp. 77 – 100 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Brazzel , J. R. , and B. G. Hightower. 1960 . A seasonal study of diapause, reproductive activity and seasonal tolerance to insecticides in the boll weevil . J. Econ. Entomol . 53 : 41 – 46 . Google Scholar Crossref Search ADS WorldCat Brazzel , J. R. , and L. D. Newsom. 1959 . Diapause in Anthonomus grandis Boh . J. Econ. Entomol . 52 : 603 – 611 . Google Scholar Crossref Search ADS WorldCat Brazzel , J. R. , T. B. Davich, and L. D. Harris. 1961 . A new approach to boll weevil control . J. Econ. Entomol . 54 : 723 – 730 . Google Scholar Crossref Search ADS WorldCat Burke , H. R. , W. E. Clark, J. R. Cate, and P. A. Fryxell. 1986 . Origin and dispersal of the boll weevil . Bull. Entomol. Soc. Am . 32 : 228 – 238 . Google Scholar OpenURL Placeholder Text WorldCat Carter , F. L. , T. C. Nelson, A. G. Jordan, and J. R. Smith. 2001 . US cotton declares war on the boll weevil, pp. 7 – 24 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Chesnut , T. L. , and W. H. Cross. 1971 . Arthropod parasites of the boll weevil, Anthonomus grandis: 2. Comparisons of their importance in the United States over a period of thirty-eight years . Ann. Entomol. Soc. Am . 64 : 549 – 557 . Google Scholar Crossref Search ADS WorldCat Clark , D. O . 2001 . Boll weevil establishment and eradication in California, pp. 235 – 244 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Coad , B. R. , E. S. Tucker, W. B. Williams, F. F. Bondy, and R. C. Gaines. 1922 . Dispersion of the boll weevil in 1921 . USDA Dept. Circ . 210 : 1 – 3 . Google Scholar OpenURL Placeholder Text WorldCat Cross , W. H . 1973 . Biology, control, and eradication of the boll weevil . Ann. Rev. Entomol . 18 : 17 – 46 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , and T. L. Chesnut. 1971 . Arthropod parasites of the boll weevil, Anthonomus grandis: 1. An annotated list . Ann. Entomol. Soc. Am . 64 : 516 – 527 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , and W. L. McGovern. 1969 . New parasites, Zatropis perdubius and Megaselia aletiae, of the boll weevil, Anthonomus grandis . Ann. Entomol. Soc. Am . 62 : 674 – 674 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , D. D. Hardee, F. Nichols, H. C. Mitchell, E. B. Mitchell, P. M. Huddleston, and J. H. Tumlinson. 1969a . Attraction of female boll weevils to traps baited with males or extracts of males . J. Econ. Entomol . 62 : 154 – 161 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , W. L. McGovern, and H. C. Mitchell. 1969b . Biology of Bracon kirkpatricki and field releases of the parasite for control of the boll weevil . J. Econ. Entomol . 62 : 448 – 454 . Google Scholar Crossref Search ADS WorldCat Cross , W. H. , M. J. Lukefahr, P. A. Fryxell, and H. R. Burke. 1975 . Host plants of the boll weevil . Environ. Entomol . 4 : 19 – 26 . Google Scholar Crossref Search ADS WorldCat Davich , T. B. , and D. A. Lindquist. 1962 . Exploratory studies on gamma radiation for sterilization of the boll weevil . J. Econ. Entomol . 55 : 164 – 167 . Google Scholar Crossref Search ADS WorldCat Fye , R. E . 1968 . The thurberia weevil in Arizona . J. Econ. Entomol . 61 : 1264 – 1268 . Google Scholar Crossref Search ADS WorldCat Fye , R. E. , C. L. Cole, F. C. Tingle, A. Stoner, D. F. Martin, and L. F. Curl. 1968 . A reproductive-diapause control program for the boll weevil in the presidio, Texas-Ojinaga, Chihuahua Area, 1965–67 . J. Econ. Entomol . 61 : 1660 – 1666 . Google Scholar Crossref Search ADS WorldCat Gast , R. T. , and T. B. Davich. 1966 . Boll weevils, pp. 405 – 418 . In Insect colonization and mass production . Academic Press , New York, NY, USA . Google Scholar Crossref Search ADS Google Preview WorldCat COPAC Giesen , J. C . 2011 . “The Herald of Prosperity”: tracing the boll weevil myth in Alabama . Agric. Hist . 85 : 24 – 49 . Google Scholar Crossref Search ADS PubMed WorldCat Haney , P. B . 2001 . The cotton boll weevil in the United States: impact on cotton production and the people of the cotton belt, pp. 7 – 24 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Hardee , D. D. , N. M. Wilson, E. B. Mitchell, and P. M. Huddleston. 1971 . Factors affecting activity of grandlure, the pheromone of the boll weevil, in laboratory bioassays . J. Econ. Entomol . 64 : 1454 – 1456 . Google Scholar Crossref Search ADS WorldCat Howard , L. O . 1894 . A new cotton insect in Texas . Insect Life , Washington . 273 . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Kuester , A. P. , R. J. Jones, T. W. Sappington, K. S. Kim, N. B. Barr, R. L. Roehrdanz, P. Senechal, J. D. Nason. 2012 . Population structure and genetic diversity of the boll weevil (Coleoptera: Curculionidae) on Gossypium in North America . Ann. Entomol. Soc. Am . 105 : 902 – 916 . Google Scholar Crossref Search ADS WorldCat Lange , F. , A. L. Olmstead, and P. W. Rhode. 2009 . The impact of the boll weevil, 1892–1932 . J. Econ. Hist . 69 : 685 – 718 . Google Scholar Crossref Search ADS WorldCat Lincoln , C . 1969 . The effect of agricultural practices on insect habitats in a typical delta community . Tall Timbers Conf Ecol Anim Contr Habitat Manage Proc. Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Lloyd , E. P. , F. C. Tingle, J. R. McCoy, and T. B. Davich. 1966 . The reproduction-diapause approach to population control of the boll weevil . J. Econ. Entomol . 59 : 813 – 816 . Google Scholar Crossref Search ADS WorldCat Lloyd , E. P. , F. C. Tingle, M. E. Merkl, E. C. Burt, D. B. Smith, and T. B. Davich. 1967 . Comparison of three rates of application of ultra-low-volume azinphosmethyl in a reproduction-diapause control program against the boll weevil . J. Econ. Entomol . 60 : 1696 – 1699 . Google Scholar Crossref Search ADS WorldCat Lloyd , E. P. , M. E. Merkl, F. C. Tingle, W. P. Scott, D. D. Hardee, and T. B. Davich. 1972 . Evaluation of male-baited traps for control of boll weevils following a reproduction-diapause program in Monroe County, Mississippi . J. Econ. Entomol . 65 : 552 – 555 . Google Scholar Crossref Search ADS WorldCat McKibben , G. H. , E. J. Villavaso, W. L. McGovern, and B. Grefenstette. 2001 . United States department of agriculture – research support, methods development and program implementation, pp. 101 – 136 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher, Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Neal , C. R. , and L. Antilla. 2001 . Boll weevil establishment and eradication in Arizona and Northwest Mexico, pp. 213 – 224 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Parencia , C. R . 1959 . Comparative yields of cotton in treated and untreated plots in insect-control experiments in central texas, 1939–1958 . J. Econ. Entomol . 52 : 757 – 758 . Google Scholar Crossref Search ADS WorldCat Parencia , C. R. , and C. B. Cowan. 1960 . Increased tolerance of the boll weevil and cotton fleahopper to some chlorinated hydrocarbon insecticides in central texas in 1958 . J. Econ. Entomol . 53 : 52 – 56 . Google Scholar Crossref Search ADS WorldCat Pierce , J. B. , C. A. Sutherland, L. J. Dominguez, A. B. Miller, S. S. Sanderson, and J. A. Calvani. 2001 . Boll weevil establishment and eradication in New Mexico, pp. 361 – 374 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Rainwater , C. F. , and J. C. Gaines. 1951 . Seasonal decline in the effectiveness of certain insecticides against boll weevil . J. Econ. Entomol . 44 : 971 – 974 . Google Scholar Crossref Search ADS WorldCat Ramalho , F. S. , and P. A. Wanderley. 1996 . Ecology and management of the boll weevil in South American cotton . Am. Entomol . 42 : 41 – 47 . Google Scholar Crossref Search ADS WorldCat Raszick , T. J. , C. M. Dickens, L. C. Perkin, A. E. Tessnow, C. P. Suh, R. Ruiz-Arce, T. N. Boratynski, M. R. Falco, J. S. Johnston, and G. A. Sword. 2021 . Population genomics and phylogeography of the boll weevil, Anthonomus grandis Boheman (Coleoptera: Curculionidae), in the United States, northern Mexico, and Argentina . Evol. Appl . 14 : 1778 – 1793 . Google Scholar Crossref Search ADS PubMed WorldCat Ridgway , R. L. , and H. C. Mussman. 2001 . Integrating science and stakeholder inputs - the pivotal years, pp. 55 – 76 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Roehrdanz , R. L . 2001 . Genetic differentiation of Southeastern boll weevil and thurberia weevil populations of Anthonomus grandis (Coleoptera: Curculionidae) Using Mitochondrial DNA . Ann. Entomol. Soc. Am . 94 : 928 – 935 . Google Scholar Crossref Search ADS WorldCat Roussel , J. S. , and D. F. Clower. 1957 . Resistance to the chlorinated hydrocarbon insecticides in the boll weevil . J. Econ. Entomol . 50 : 463 – 468 . Google Scholar Crossref Search ADS WorldCat Rummel , D. R. , and P. L. Adkisson. 1971 . A two-phased control program designed for maximum suppression of the boll weevil in the high and rolling plains of texas . J. Econ. Entomol . 64 : 919 – 922 . Google Scholar Crossref Search ADS WorldCat Stavinoha , K. D. , and L. A. Woodward. 2001 . Texas boll weevil history, pp. 451 – 502 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Sterling , W. L . 1978 . Fortuitous biological suppression of the boll weevil by the red imported fire ant . Environ. Entomol . 7 : 564 – 568 . Google Scholar Crossref Search ADS WorldCat Tate , J. R . 2001 . Cotton production and the boll weevil in Virginia, pp. 503 – 508 . In W. A. Dickerson, A. L. Brashear, J. T. Brumley, F. L. Carter, W. J. Grefenstette and F. A. Harris (eds.), Boll weevil eradication in the United States through 1999, the cotton foundation reference book series . The Cotton Foundation Publisher , Memphis, TN, USA . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC Tumlinson , J. H. , D. D. Hardee, R. C. Gueldner, A. C. Thompson, P. A. Hedin, and J. P. Minyard. 1969 . Sex pheromones produced by male boll weevil: isolation, identification, and synthesis . Science . 166 : 1010 – 1012 . Google Scholar Crossref Search ADS PubMed WorldCat Vanderzant , E. S. , and T. B. Davich. 1958 . Laboratory rearing of the boll weevil: a satisfactory larval diet and oviposition studies . J. Econ. Entomol . 51 ( 3 ): 288 – 291 . Google Scholar Crossref Search ADS WorldCat Warner , R. E . 1966 . Taxonomy of the Subspecies of Anthonomus grandis (Coleoptera: Curculionidae) . Ann. Entomol. Soc. Am . 59 : 1073 – 1088 . Google Scholar Crossref Search ADS WorldCat Warner , R. E. , and C. E. Smith , Jr. 1968 . Boll weevil found in pre-columbian cotton from Mexico . Science . 162 : 911 – 912 . Google Scholar Crossref Search ADS PubMed WorldCat Yang , C. , C. P. C. Suh, and J. K. Westbrook. 2017 . Early identification of cotton fields using mosaicked aerial multispectral imagery . J. Appl. Remote Sens . 11 : 016008 . Google Scholar Crossref Search ADS WorldCat © The Author(s) 2021. Published by Oxford University Press on behalf of Entomological Society of America. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com. This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)

Journal

Annals of the Entomological Society of AmericaOxford University Press

Published: Oct 5, 2021

There are no references for this article.