You are on page 1of 16

Impact of climate change on plant pests and diseases Climate change will affect plant pests and diseases

in the same way it affects infectious disease agents. In other words, the range of many insects will expand or change, and new combinations of pests and diseases may emerge as natural ecosystems respond to altered temperature and precipitation profiles. Any increase in the frequency or severity of extreme weather events, including droughts, heat waves, windstorms, or floods, could also disrupt the predator-prey relationships that normally keep pest populations in check. An explosion of the rodent population that damaged the grain crop in Zimbabwe in 1994, after 6 years of drought had eliminated many rodent predators, shows how altered climate conditions can intensify pest problems. The effect of climate on pests may add to the effect of other factors such as the overuse of pesticides and the loss of biodiversity that already contribute to plant pest and disease outbreaks (McMichael et al,1996). The ingenuity of farmers, breeders, and agricultural engineers, and the natural resilience of biological systems, will help buffer many of the negative effects of climate change on agriculture. However, experts believe that over the longer term, the accumulated stresses of sustained climate change stand a good chance of disrupting agro-ecosystems and reducing global food productivity. The regions thought to be most vulnerable to productivity declines are semiarid and arid areas where rain-fed, non irrigated agriculture predominates. Unfortunately, many of these areas as in sub-Saharan Africa, South and East Asia, and on some Pacific islands are already hard-pressed agriculturally and suffer from high rates of malnutrition. In Senegal, for example, one study predicts a 30-percent yield decline with a 4?C rise in temperatures and no change in rainfall from current levels. The effect of this kind of agricultural decline on local food security would be severe. The negative effects of climate change on agriculture in poor countries could put an additional 40 to 300 million people at risk of hunger by 2060 (McMichael et al, 1996). Perhaps the greatest long-term danger to human health from climate change will be the disruption of natural ecosystems, which provide an array of services that ultimately support human health. Biotic systems whether in forests, rangelands, aquatic environments, or elsewhere provide food, materials, and medicines; store and release fresh water; absorb and detoxify wastes; and satisfy human needs for recreation and wilderness. They are also intimately involved in sustaining the genetic basis of agriculture. These systems will likely undergo major reorganization as global temperatures rise and rainfall patterns change more rapidly than they have in the past 10,000 years. Rough estimates of the effects of a doubling of atmospheric CO2 levels show a major redistribution of Earths vegetation. As much as one third to one half of all plant communities and the animals that depend on them might shift in response to changing ecological conditions.

In general, low-latitude areas are most at risk of having decreased crop yields. In contrast, mid- and high-latitude areas could generally, although not in all locations, see increases in crop yields for temperature increases of up to 1-3C. Taken together, there is low to medium confidence that global agricultural production could increase up to approximately 3C of warming. For temperature increases beyond 1-3C, yields of many crops in temperate regions are projected to decline. As a result, beyond 3C warming, global production would decline because of climate change and the decline would continue as GMT increases. Most studies on global agriculture have not yet incorporated a number of critical factors, including changes in extreme events or the spread of pests and diseases. In addition, they have not considered the development of specific practices or technologies to aid adaptation. Some studies (Rosenzweig et al., 2001; FAO, 2004) agree that higher temperatures and longer growing seasons could result in increased pest populations in temperate regions of Asia. CO2 enrichment and changes in temperature may also affect ecology, the evolution of weed species over time and the competitiveness of C3 v. C4 weed species (Ziska, 2003). Warmer winter temperatures would reduce winter kill, favouring the increase of insect populations. Overall temperature increases may influence crop pathogen interactions by speeding up pathogen growth rates which increases reproductive generations per crop cycle, by decreasing pathogen mortality due to warmer winter temperatures, and by making the crop more vulnerable. Climate change, as well as changing pest and disease patterns, will likely affect how food production systems perform in the future. This will have a direct influence on food security and poverty levels, particularly in countries with a high dependency on agriculture. In many cases, the impact will likely be felt directly by the rural poor, as they are often closely linked to direct food systems outcomes for their survival and are less able to substitute losses through food purchases. The urban poor are also likely to be affected negatively by an increase in food prices that may result from declining food production. Studies suggest that temperature increases may extend the geographic range of some insect pests currently limited by temperature (Poter,1990) The European Corn Borer is a major pest of grain maize in many parts of the world. It is multivoltine (multigenerational) and, depending on climatic conditions, can produce up to four generations per year. Using degree-day (thermal) requirements, the potential distribution of the European Corn Borer in Europe has been mapped under present (1951-80) temperatures. With a 1deg.C increase in temperature a northward shift in distribution of between 165 and 500 km is indicated for all generations. In addition to favourable climatic conditions the distribution of any pest is dependent on the availability of a host plant. The potential limit of grain maize cultivation is also likely to shift northwards with increasing temperatures providing suitable conditions for the European Corn Borer. This example serves to highlight the need to examine crop-pest interactions in any climate impact assessment. Under a warmer climate at mid-latitudes there would be an increase in the overwintering range and population density of a number of important agricultural pests, such as the potato leafhopper which is a serious pest of soybeans and other crops in the USA (Smith, and Tirpak, 1989). Assuming planting dates did not change, warmer temperatures would

lead to invasions earlier in the growing season (i.e. at more susceptible stages of plant development) and probably lead to greater damage to crops. In the US Corn Belt increased damage to soybeans is also expected due to earlier infestation by the corn earworm, which could result in serious economic losses. One of the major threats of climatic change is the establishment of "new" or migrant pests as climatic conditions become more favorable to them. In New Zealand, for example, the swarming of locusts in the North Island in recent years may be an indication of a more widespread problem in the future (Messenger, 1989). In cool temperate regions, where insect pests and diseases are not generally serious at present, damage is likely to increase under warmer conditions. In Iceland, for example, potato blight currently does little damage to potato crops, being limited by the low summer temperatures. However, under a 2 x CO2 climate that may be 4deg.C warmer than at present, crop losses to disease may increase to 15 per cent (Bergthorssonet al,1988). Most agricultural diseases have greater potential to reach severe levels under warmer conditions. Fungal and bacterial pathogens are also likely to increase in severity in areas where precipitation increases (Beresford and Fullerton, 1989). Under warmer and more humid conditions cereals would be more prone to diseases such as Septoria. In addition, increases in population levels of disease vectors could lead to increased epidemics of the diseases they carry. To illustrate, increases in infestations of the Bird Cherry aphid (Rhopalosiphum padi) or Grain aphid (Sitobian avenae) could lead to increased incidence of Barley Yellow Dwarf virus in cereals. Climate change implications for insect pests 1. Climate change is likely to alter the balance between insect pests, their natural enemies and their hosts; predictions of the impact of climate change on insect damage to UK forests are therefore difficult to make. 2. One of the most important effects of climate change will be to alter the synchrony between host and insect pest development, particularly in spring, but also in autumn; the predicted rise in temperature will also generally favour insect development and winter survival, although there will be some exceptions. 3. The green spruce aphid is one example of an insect that is likely to benefit from the increase in winter survival, leading to more intense and frequent tree defoliation. A decline in the productivity of Sitka spruce might therefore be expected. 4. Modelling work suggests that under a warmer climate, exotic pests such as the southern pine beetle could establish populations in Europe, and that climatic warming could make UK forests susceptible to damage; other bark beetles such as Ips typographus, which is present in some parts of Europe, but not the UK, could become a serious problem.

5. Rising atmospheric CO2 concentrations may lead to a decline in food quality for plant-feeding insects, as a result of reduced foliar nitrogen levels. 6. The planting of exotic tree species may exacerbate the beneficial effects of climate change on insect pests, as the natural predatory fauna may not be present to limit population growth. 7. Changes have already been observed in the distribution of native European butterfly populations, with northern ranges extended and southern ranges reduced. The same is likely to be the case for forest insect pests. 8. The combined effects of increased global trafficking of timber and wood products and climate change are likely to result in exotic pests such as Asian longhorn beetle becoming more prevalent; it is therefore essential that we remain vigilant in reporting new pests and altered patterns of damage. Climate and the physiology of animals and plants Higher levels of carbon dioxide could stimulate the growth of some weed species, especially summer-active weeds in higher rainfall zones. However, the decline in rainfall predicted for southern Australia may counteract this. Woody weeds will benefit from increased carbon dioxide more than grasses. Many plant species respond to accumulated day degreesthe cumulative sum of daily temperaturesto read the season and trigger critical development stages such as stem elongation and flowering. Warmer temperatures will accelerate the rate at which day degrees accumulate, so the life cycles of some plant species may accelerate. Because plants are host to many pest animals, the life cycle of some pest species, such as Red Legged Earth Mite, aphids and rabbits, will respond to their plant hosts and change their feeding and reproductive patterns accordingly. Warmer temperatures will directly affect the ability of animals to maintain their body temperature and avoid heat stress. Increased levels of carbon dioxide can affect the carbon-to-nitrogen ratio of plant material, thereby reducing the nitrogen available to plant-eating animals and insects. Insects that need lower temperatures to activate dormancy may have shorter overwintering periods. Warmer temperatures may reduce the production of dew which is an important source of moisture for many insects and smaller vertebrate species. Pest control strategies will change Across farming regions in southern Australia, the winter growing season is expected to shorten. This will have implications for weeds and for plant-eating pests such as rabbits. Warmer winter temperatures will accelerate the development of plant species that use accumulated day degrees to trigger developmental stages (e.g. tillering, flowering). As a result, pastures, crops and weeds are likely to mature and start to decay earlier. Accelerated plant development has important implications for crop and pasture protection and for weed control. For example, producers may need to adjust the timing of the control of Red Legged Earth Mite as the mites life cycle responds to climate change. Similarly, the timing of pasture topping may need adjusting to match changes in season length and plant development rates.

In lower-rainfall regions, a shorter winter growing season will mean longer periods of dry feed over summer. This may reduce the breeding window for rabbits and feral pigs, and could increase the effectiveness of some control programs. Feed and water shortages in extreme seasons may increase competition between kangaroos and domestic livestock; kangaroos may be better able to take advantage of changed conditions and become dominant. At the same time, feral goats and camels may have a competitive advantage over native species, posing a threat to conservation but offering a commercial harvest opportunity. Baiting to control pest animals such as foxes may be more effective when food is scarce. As insect breeding cycles adjust to changes in season timing and length, producers may need to change the timing of insect control using, for example, the sterile insect technique and bait spraying. Pest populations will change Projected warming will help some pest species to survive winter and will accelerate development of summer-active species. While projections for rainfall are less certain than those for temperature, summer rainfall, particularly in southern Australia, is not predicted to decline as much as winter rainfall. So summer weeds, in particular, are likely to do well. Blowfly populations could also increase because flies breed better and grow faster in warmer conditions. Higher temperatures may reduce the reproductive potential of a number of important agricultural pests. Rabbits need cool temperatures to breed and when the temperature is above 29oC females cannot produce enough milk for their young. Lice populations are also expected to decline in hotter environments. Adult lice can move around an animal to maintain an optimal temperature of 37oC, but lice eggs cannot survive extreme temperatures. Germination of winter annual weeds may be affected by reduced winter rainfall. Pests will migrate In any particular location, climate change may not mean more pest animals and weeds, but it could mean new pest animals and weeds. The range of pests will generally move southwards and shift to higher altitudes as a result of warming trends. Temperate weeds are expected to be displaced southwards. During wetter years, southern Australia will have the potential to host traditionally semi-tropical weeds and parasites, such as lantana and cattle ticks. Temperature-sensitive species such as blackberry are already invading higher altitudes, and frost-intolerant species such as rubber vine and Chromolaena are predicted to move southwards. As the potential range of weeds expands with climate change, the actual range may lag while the weed species adapt to the higher temperatures. Weeds with efficient seeddispersal systems (wind, water, birds) will invade more quickly than weeds that rely on vegetative dispersal. An increase in extreme events, such as cyclones, storms and associated floods, may increase the dispersal of weed species that rely on wind and water to move seeds or pollen. Cyclones and storms may also help the spread of insects within Australia and facilitate exotic species incursions from neighbouring islands. Habitats disturbed by extreme events such as drought leave empty niches which pest animals and weeds can colonise. Pests often recover from extreme climatic events faster than other species. Weed Impact

Blackberry Expected to retreat southwards and to higher altitudes because it is sensitive to higher temperatures and drought. Chilean needle grass Expected to increase its range because it is highly invasive (longlived, seed dispersed by wind and water) and drought tolerant. Gorse Expected to retreat southwards because it is drought sensitive. Lantana Expected to continue its move southwards into high-rainfall zones of northern New South Wales. Mesquite Some risk that it may move into lower-rainfall areas because it is very drought tolerant. Parthenium Not suited to winter-dominant rainfall areas. May move into summer-dominant, higher-rainfall (>500 mm) regions. Serrated tussock Expected to retreat southwards and to higher altitudes because it is sensitive to higher temperatures. As a drought-tolerant plant, it should become more invasive in areas where temperature allows. Prickly acacia Expected to move southwards and into arid areas. Effectiveness of pest control will change Climate change will impact the effectiveness of pesticides. Weeds that are under moisture stress can respond by thickening their leaf cuticles, slowing down vegetative growth and flowering rapidly. Drought-stressed weeds are more difficult to control with postemergent herbicides than plants that are actively growing; for example, systemic herbicides that are translocated within the weed need active plant growth to be effective. Pre-emergent herbicides or herbicides absorbed by plant roots need soil moisture and actively growing roots to reach their target species. Drying winter and spring rainfall trends have the potential to reduce the effectiveness of pre-emergent herbicides such as triazines or atrazine. New environmental weeds will emerge Climate change will provide the opportunity for environmental weeds to invade new ecosystems. Environmental weeds not associated with agriculture have the potential to significantly reduce the biodiversity value of on-farm native vegetation and riparian areas. They are hard to control once established, because it is difficult to use traditional control methods in native bush and riparian areas. Drying soil conditions and lower incidence of frost could increase the range of bitou bush and bone seed, which are significant environmental weeds. Producers should monitor and report new weed outbreaks in native vegetation to facilitate control of their spread. Opportunities for producers Producers are likely to have time to adapt their pest management strategies to climate change. Continued strategic implementation of pest control measures coordinated across jurisdictions will help protect agriculture from the full impact of climate change over the medium term. Climate change will offer opportunities to control traditional pests; for example, producers can learn from strategies used in the north to control southward-moving pests. Producers can take advantage of the natural stress conditions that traditional pest species will face, by reducing population numbers and host species whenever economically feasible and encouraging the adaptation of desirable species. Climate change may help to control some pests, such as rabbits, blackberries, serrated tussock, gorse and lice. Rabbits may struggle with the longer periods of dry feed, higher temperatures and expected increased effectiveness of calicivirus under drier conditions. Adapting to the changing climate

Australian producers are already adapting to a changing climate. They should continue to adapt by being well prepared to manage the condition of their land, including seasonal outbreaks of pests and weeds. Understanding that current pest animal and weed species may decline and that new pests may become a threat is a key part of preparing for and adapting to climate change. Working to prevent the establishment of new pest animal and weed species must be a priority for all landholders. Climate change may provide an opportunity for vigilant landholders to tackle new pests before they become established. To maximise the opportunities presented by climate change, pest management practices must adapt with the biology and ecology of the pests and producers need to understand Pest and disease shifts Pest species are also moving poleward and upward. Over the past 32 years, the pine processionary moth (Thaumetopoea pityocampa) has expanded 87 km at its northern range boundary in France and 110230 m at its upper altitudinal boundary in Italy. Laboratory and field experiments have linked the feeding behavior and survival of this moth to minimum nighttime temperatures, and its expansion has been associated with warmer winters. In the Rocky Mountain range of the United States, mountain pine beetle (Dendroctonus ponderosae) has responded to warmer temperatures by altering its life cycle. It now only takes one year per generation rather than its previous two years, allowing large increases in population abundances, which, in turn, have increased incidences of a fungus they transmit (pine blister rust, Cronartium ribicola) (Logan et al. 2003). Increased abundance of a nemotode parasite has also occurred as its life cycle shortened in response to warming trends. This has had associated negative impacts on its wild musk oxen host, causing decreased survival and fecundity. In a single year (1991), the oyster parasite Perkinsus marinus extended its range northward from Chesapeake Bay to Mainea 500 km shift. Censuses from 1949 to 1990 showed a stable distribution of the parasite from the Gulf of Mexico to its northern boundary at Chesapeake Bay. The rapid expansion in 1991 has been linked to aboveaverage winter temperatures rather than human-driven introduction or genetic change. A kidney disease has been implicated in low-elevation trout declines in Switzerland. High mortality from infection occurs above 1516C, and water temperatures have risen in recent decades. High infection rates (27% of fish at 73% of sites) at sites below 400 m have been associated with a 67% decline in catch; mid-elevation sites had lower disease incidence and only moderate declines in catch; and the highest sites (8003029 m) had no disease present and relatively stable catch rates (Hari et al. 2006). Changes in the wild also affect human disease incidence and transmission through alterations in disease ecology and in distributions of their wild vectors (Parmesan and Martens 2006). For example, in Sweden, researchers have documented marked increases in abundances of the disease-transmitting tick Ixodes ricinus along its northernmost range limit (LindgrenandGustafson 2001). Between the early 1980s and 1994, numbers of ticks found on domestic cats and dogs increased by 22%44% along the ticks northern range boundary across central Sweden. In the same time period, this region had a marked decrease in the number of extremely cold days (<12C) in winter and a marked increase in warm days (>10C) during the spring, summer, and fall. Previous studies on temperature developmental and activity thresholds indicated the observed warmer temperatures cause decreased tick mortality and longer growing seasons (Lindgren and Gustafson 2001).

How might climate change affect insect pest outbreaks? It is difficult to predict the impacts of climate change on forest insect pests because of the complexity of the interactions between insects and trees. This overall response is dependent on the impacts of climate change on the insect- tree host-natural enemy relationship. However, some generalised predictions can be made, based on current pest distributions and the severity of insect outbreaks in individual regions. New (alien or exotic) insect species that may present a future threat to woodland are also assessed, although their introduction is as likely to be driven by global trade as by climate change. The analysis is based on expert judgement, and is very much a 'vision of the future' and not 'hard science'. It should be placed in the context of indicating those insect pests for which there is an ongoing and future need for monitoring, and not as a hard-and-fast prediction of future impacts. Elatobium abietinum (green spruce aphid). Populations are largely controlled by low winter temperatures and over-wintering populations are therefore thought likely to increase with climate change. This effect of temperature on population dynamics has been clearly demonstrated and serious outbreaks are predicted to become more commonplace. Research indicates that a serious outbreak reduces increment in the year of damage by 20-30%. In addition, a higher frequency of outbreaks could have more profound effects on site productivity. Hylobius abietis. A serious and widespread problem during restocking of conifer plantations. The weevil has a large climatic range (across most of Europe), although activity increases to some extent with temperature. Climatic warming might thus be expected to increase its impact, although the use of small felling coups which is now common standard practice, particularly in CCF systems, may act to moderate any increase in activity. Dendroctonus micans (great spruce bark beetle). Infestations are sporadic, but excessive bark damage leads to tree death, especially in Sitka spruce. The species is likely to benefit from an increased frequency of summer drought and climatic warming, although its specific predator (Rhizophagus grandis) might benefit to a greater extent, thus potentially reducing the impact of Dendroctonus. Ips acuminatus (and others: bark beetle on pine trees). Reports of serious damage in the UK are not common. It is largely a secondary pest, but primary infestations have been reported, particularly at the young thicket stage on light soils. Primary attacks are linked to drought stress, so again, climate change may worsen the situation in some areas. Population build-up also occurs in areas of wind-blow, which

may therefore present a future climate change impact should storm damage become more commonplace. Ips cembrae (large larch bark beetle). Its distribution is currently limited to areas in Scotland and northern England where it is a primary agent of tree death. If populations were established in other areas, Ips cembrae could cause serious damage in stands subjected to the increased frequency of summer drought that is predicted for much of England. Pissodes castaneas (small-banded pine weavil). Infestations have been periodic and heavy, and characterised by top-death of affected trees. Drought stress is seen as a trigger for infestation; climate change is therefore expected to make infestations more frequent and severe. Rhyacionia buoliana (pine shoot moth). Serious infestations occur when trees are under stress. It is thought likely that its impact will increase with climate change. Agrilus pannonicus (buprestid beetle). Associated with oak decline, but it is not certain whether the relationship is causative, and there is also conjecture as to whether it is a primary or secondary pest it is seen as a primary agent of decline in Germany, but only as a secondary agent in the later stages of decline in the UK . At present, its distribution is restricted to southern England. Climatic warming may extend its range. Platypus cylindicus (oak beetle). A secondary pest affecting the value of felled broadleaf timber, particularly oak. It was a red data-book species in the early 1980s, but became a serious pest following the 1987 storms as a result of large quantities of lying timber. Is only likely to be a problem if there are widespread losses for other reasons (drought or storm). Longhorn beetles. Only infest dead trees at present, but may attack living trees under conditions of extreme drought. They are relatively common in old timber. Tortrix and winter moth At present there is no link between climate and damaging episodes. However, if synchrony were to break down between the egg hatch and canopy development in the spring (which it has not to date), defoliating episodes may become less common.

Cameraria ohridella (horse chestnut leaf miner). This insect has spread rapidly across Europe from Macedonia, in northern Greece, and is now established in London and other towns in south east England. It is thought that the speed of spread may be related to recent climate change. A Forestry Commission Exotic Pest Alert on the moth has recently been published. It is likely that the range of the leaf miner will expand. The long-term impact of the moth and methods of control are currently under investigation. Sesia apiformis (hornet moth). This species is a stem borer of poplar, and its prevalence may be linked to climate. It is a secondary pest, and only presents a problem when trees are stressed by summer drought. New alien/exotic pests include the following, and vigilance should be maintained, particularly regarding the importation of timber and wood products. Anoplophora glabripennis (Asian longhorn beetle). Populations could establish in warmer coastal areas if the beetle was introduced, and they would benefit from climatic warming. Amenity and street trees are particularly susceptible. Ips typographus (eight-toothed spruce bark beetle). Pheromone traps in ports have recently recorded higher numbers than in previous years. If predictions of increased soil moisture deficits are borne out, drought-stressed trees could provide a resource from which wider outbreaks might develop. As with Ips acuminatus, populations build up in areas of significant wind-blow. In general, it is a reasonable assumption that many insect pests have the potential to become more damaging as a result of climate change. Most of these predictions are driven through expectations that trees across much of Britain will become more drought stressed during the summer months, and that development of pests will accelerate, possibly leading to higher population growth. At the same time, it is likely that the pests natural enemies will also benefit, so it is, to some extent, unclear as to what the overall effects will be. Climate change could disrupt natural pest control Climate change could upset the balance between insect crop pests and the 'natural enemies' that control their numbers, say researchers. This might make pest outbreaks more frequent and severe, particularly in the tropics where the climate is naturally more consistent. Global warming is expected to make regional climates more varied and unpredictable, which could affect relationships between plant-eating insects and their enemies. The study revealed that significantly fewer parasitoids were produced in years when rainfall was most variable. This could be because the parasitoids use cues such as changes

in local climate to determine the best time for laying their eggs. If so, unpredictable rains might disrupt the parasitoids' ability to 'track' their caterpillar hosts. "The wasps use the start of the rains as a cue to hatch out of their cocoons and look for a caterpillar to lay their eggs on. "If the rains are late, they will emerge late, and miss the narrow window of time when their host is most vulnerable." "Other agents, such as predators, diseases and fungal pathogens, might be affected differently by climate change," he said. The impact of climatic change on horticultural pests The climate has profound effects on populations of invertebrate pests (e.g. insects, mites, slugs), affecting their development, reproduction and dispersal. The rate at which most pests develop is dependent on temperature and every species has a particular threshold temperature above which development can occur, and below which development ceases. As temperatures rise, some pest species may be able to complete more generations in a year. This effect may be most noticeable in insects with short life-cycles such as aphids and the diamond-back moth. On the other hand, the temporary exposure of populations to extremely high temperatures may delay the development of surviving individuals and thus delay the subsequent generation. The increased frequency of heat waves will undoubtedly be deleterious to the survival of some pests such as carrot fly, which is particularly sensitive to a combination of high temperature and dry soil conditions during the egg and early larval stages. In contrast, an increase in winter temperatures will improve survival of some pests. Notable examples are the species of aphid that spend the winter as adults and nymphs (peach-potato aphid, cabbage aphid, potato aphid) and other pests. Too much water can be devastating for some pests. Raindrops can physically dislodge them from their host plant and behaviour patterns can be disrupted. Some pest infestations are suppressed by periods of rainfall, either because of physical effects or because high humidity leads to outbreaks of fungal disease, often observed amongst aphids on lettuce and brassica crops. Some pest insects appear to prosper during periods of drought and cutworms (caterpillars of the turnip moth) appear to be one such example. Some species are particularly well adapted to moving long distances in the air and these include several species of moth, such as the diamond-back moth and silver Y moth. Wind direction, and high altitude wind in particular, is important to the flights of migratory pests. Climate change may affect our ability to control pests. For example, high temperature is reported to reduce the effectiveness of some pesticides. Humidity levels can also modify their efficacy, as can the timing and amount of rain following their application. On a simpler level, rain can affect growers ability to apply the pesticide at the time of most need, possibly an increasingly likely scenario. If pests are able to complete more

generations in a season then this may lead to greater pesticide use, which in turn may lead to the more rapid development of pesticide resistance. Strategies to avoid the development of resistance need to be planned in advance and rely on the availability of a range of pesticides and/or alternative control methods. The natural enemies of pests will have their own climate optima, although not necessarily the same as their hosts. It is a pertinent to ask whether climate change will affect natural enemies to the same extent, or in the same direction, as the pests they control. The advent of milder winters and warmer summers, more typical currently of other parts of Europe, has implications for the survival and reproduction of new pests. These may arrive as migrants, or on imported plant material. The Colorado beetle and the western corn root worm (Diabrotica virgifera virgifera) are good examples. Obviously, there is considerable uncertainty attached to climate change predictions and this is true also of likely changes in pest pressure in the future. Some pests will inevitably become a more serious problem, whilst others may decline. Whatever the long term consequences of climate change, pest numbers will undoubtedly continue to fluctuate from season-to-season, depending on the particular combination of weather conditions that occurs each year. Predicting the impact of climate change on insects and diseases is a very complex exercise and one that involves a great deal of modeling. There have been some rather spectacular documented cases of how climate can influence insect populations. One good example getting press right now is the mountain pine beetle situation on the West Coast. Climate plays a critical role in the ability of insects to overwinter, on the geographic distribution or ranges of insects, on the number of generations and on their abundance in agricultural systems. Where an insect chooses to overwinter may affect its survival. For example, some insects overwinter near the soil surface. In this case, snow cover can provide an excellent insulation, and if this is reduced, these insects might not survive a very cold winter. On the other hand, an increase in temperature and a longer growing season could result in an extension of ranges of plants and animals, with new insects being able to establish themselves in a region where they were previously unable to survive. So new pests would require the development of monitoring programs, the establishment of action thresholds and the availability of new pest control products and strategies for IPM. All of these things take time to develop. Similarly, an increase in the number of generations can translate into the need for additional controls and would present challenges to resistance management. Some insects that are normally secondary pests might also become more serious. Altered wind patterns could affect the long distance movement of some species that do not overwinter are picked up and moved by weather fronts. Since many insects are vectors of plant diseases, subtle changes in climate could affect the incidence of transmission. There are still many unknowns in the climate change

equation. The general consensus is that extremes of temperature will become more common. With an overall warming trend, insect woes just might increase. Pest management strategies in agriculture and forestry will require adjustment. References Aerts, M., P. Cockrell, G. Nuessly, R. Raid, T. Schueneman, and D. Seal. 1999. Crop Profile for Corn (Sweet) in Florida. http://www.ipmcenters.org/CropProfiles/docs/FLcornsweet.html Andrew, N.R. and L. Hughes. 2005. Diversity and assemblage structure of phytophagous agricultural insects. Ann Rev Entomol 48: 261-281. Awmack, C.S., C.M. Woodcock and R. Harrington. 1997. Climate change may increase vulnerability of aphids to natural enemies. Ecological Entomology. 22:366-368. Bale, J.S. G.J. Masters, I.D. Hodkinson, C. Awmack, T.M. Bezemer, V.K. Brown, J. Butterfield, A. Buse, J.C. Coulson, J. Farrar, J.E.G. Good, R. Harrington, S. Hartley, T.H. Baniecki, J. and M. Dabaan. 2000a. Crop Profile for Corn (sweet) in West Virginia. http://www.ipmcenters.org/CropProfiles/docs/wvcorn-sweet.html Baniecki, J. and M. Dabaan. 2000b. Crop Profile for Potatoes in West Virginia. http://www.ipmcenters.org/CropProfiles/docs/WVpotatoes.html Bergthorsson, P., Bjornsson, H., Dyrmundsson, O., Gudmundsson B., Helgadottir, A., and Jonmundsson, J.V.1988. "The effects of climatic variations on agriculture in Iceland", in Parry, M.L., Carter, T.R., and Konijn, N.T., (eds), The Impact of Climatic Variations onAgriculture, Volume 1, Assessments in Cool Temperate and Cold Regions (Dordrecht, The Netherlands: Kluwer, 1988), pp.383-509. Hoffman, B. 1999. Crop Profile for Potatoes in Pennsylvania. http://www.ipmcenters.org/CropProfiles/docs/papotatoes.html Houghton, J.T, Meiro Filho, L.G, Callander, B.A, Harris, N, Kattenberg A, Maskell K (eds) (1996) Climate Change 1995: The Science of Climate Change. Contribution of Working Group I to the Second Assessment of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, 584pp. Hunter, M.D. 2001. Effects of elevated atmospheric carbon dioxide on insect-plant interactions. Ag. Forest. Entomol. 3:153-159. Karl TR, Nichols N, Gregory J. 1997. The coming climate. Scientific American, 276, 54 59. Kaukoranta, T. 1996. Impact of global warming on potato late blight: risk, yield loss, and control. Agric. Food Sci. Finl. 5:311-327. Kenny, G.J, Harrison, P.A, Parry, M.L. 1993. The Effect of Climate Change on Agricultural and Horticultural Potential in Europe. Research Report no. 2, January 1993. Environmental Change Unit, University of Oxford. Leemans R and Soloman A.M .1993. Modelling the potential change in yield and distribution of the earths crops under a warmed climate. Climate Research, 3, 7996. Lewis, T. 1997. Thrips as crop pests. CAB International, Cambridge: University Press. 740 pp. McMichael, A. et al., eds., Climate Change and Human Health (World Health Organization, Geneva, 1996), p. 218. Mcvean, R. and Dixon, A. F. G. 2001. The effect of plant drought-stress on populations of the pea aphid Acyrthosiphon pisum. Ecol. Entomol. 26: 440-443.

Messenger, G. 1989. "Migratory Locusts in New Zealand?", The Weta, vol. 11-2: 29 (1988). Hill, M.G., and Dymock, J.J., Impact of Climate Change: AgriculturallHorticultural Systems (DSlR Entomology Division, submission to New Zealand Climate Programme, Department of Scientific and Industrial Research, Wellington, New Zealand, 1989). Musser, F. P and Shelton, A. M. 2005. The influence of post-exposure temperature on the toxicity of insecticides to Ostrinia nubilalis (Lepidoptera:Crambidae). Pest Manag Sci. 61:508-510. Parry, M.L, Porter, J.H, Carter, T.R. 1990. Agriculture: climate change and its implications. Trends in Ecology and Evolution, 5, 318322. Pimentel, D. et al.,1991.Environmental and Economic Impacts of Reducing U.S. Agricultural Pesticide Use," D. Pimentel ed., in Handbook on Pest Management in Agriculture, Vol. I. (Boca Raton, FL: CRC Press), p. 679. Pimentel, D. et al., 1993. Ethical Issues Concerning Potential Global Climate Change on Food Production," Journal of Agricultural and Environmental Ethics 5., pp. 113-146. Porter, J.H 1990. IPCC, The potential impacts of climate on agriculture (Geneva and Nairobi: WMO and UNEP). Reiners, S and C. Petzoldt (eds). 2005. Integrated Crop and Pest Management Guidelines for Commercial Vegetable Production. Cornell Cooperative Extension publication #124VG http://www.nysaes.cornell.edu/recommends/ Rosenzweig, C. 1993. Recent global assessments of crop responses to climate change. In: International Crop Science I (eds Buxton DR et al.), pp. 265272. Crop Science Society of America, Madison, WI. Rosenzweig C, Parry, M.L, Fischer G, Frohberg, K. 1993. Climate change and world food supply. Research Report no. 3. Environmental Change Unit, University of Oxford, Oxford. Rosenzweig, C., A. Iglesias, X.B. Yang, P.R. Epstein and E. Chivian, 2001. Climate change and extreme weather events: implications for food production, plant diseases and pests. Global Change and Human Health, 2, 90-104 Shelton, A.M., W.R. Wilsey, and Soderlund, D.M.. 2001. Classification of insecticides and acaricides for resistance management. Dept. of Entomology, NYSAES, Geneva, NY 14456. 315-787-2352 http://www.nysaes.cornell.edu/ent/faculty/shelton/pdf/res_mgmt.pdf Smith, J.B and Tirpak, D., The Potential Effects of Global Climate Change on the United States, Report to Congress (Washington, DC: US Environmental Protection Agency, 1989). A summary is given in Adams, R.M. (and 9 others), "Global climate change and US Agriculture" Nature, vol. 345: 219-224 (1990). Sombroek WG, Gommes R (1996) The climate change agricultural conundrum. In: Global Climate Change and Agricultural Production: Direct and Indirect Effects of Changing Hydrological, Pedological and Plant Physiological processes (eds Bazzaz F, Sombroek W), pp. 114. John Wiley, Chichester.

Sparks TH, Yates TJ (1997) The effect of spring temperature on the appearance dates of British butterflies 18831993. Ecography, 20, 368374. Stivers, L. 1999a. Crop Profile for Corn (sweet) in New York. http://pestdata.ncsu.edu/cropprofiles/docs/nycorn-sweet.html Stivers, L. 1999b. Crop Profile for Potatoes in New York. http://www.ipmcenters.org/CropProfiles/docs/nypotatoes.html Vincent, C., G. Hallman, B. Panneton, and F. Fleurat-Lessard. 2003. Management of Wallin,J.R. and P.E. Waggoner. 1950. The influence of climate on the development and spread of Phytophthora infestans in artificially inoculated potato plots. Plant Dis. Reptr. Suppl. 190. pp 19-33. Watson RT, Zinyowera MC, Moss RH (1996) Climate Change 1995: Impacts, Adaptations and Mitigation of Climate Change. Contribution of Working Group II to the Second Assessment of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, 880pp. Whitney, S, J. Whalen, M. VanGessel, B. Mulrooney. 2000. Crop Profile for Corn (sweet) in Delaware. http://www.ipmcenters.org/CropProfiles/docs/DEcorn-sweet.html Yamamura, K. and K. Kiritani. 1998. A simple method to estimate the potential increase in the number of generations under global warming in temperate zones. Appl. Ent. and Zool. 33:289-298. Ziska, L.H. 2003. Evaluation of the growth response of six invasive species to past, present and future atmospheric carbon dioxide. J. Exp. Bot., 54, 395- 404

You might also like