Aphid

Last updated

Aphids
Temporal range: Permian–present
Aphid on leaf05.jpg
Scientific classification OOjs UI icon edit-ltr.svg
Domain: Eukaryota
Kingdom: Animalia
Phylum: Arthropoda
Class: Insecta
Order: Hemiptera
Suborder: Sternorrhyncha
Infraorder: Aphidomorpha
Superfamily: Aphidoidea
Geoffroy, 1762
Families

Aphids are small sap-sucking insects and members of the superfamily Aphidoidea. Common names include greenfly and blackfly, [lower-alpha 1] although individuals within a species can vary widely in color. The group includes the fluffy white woolly aphids. A typical life cycle involves flightless females giving live birth to female nymphs—who may also be already pregnant, an adaptation scientists call telescoping generations—without the involvement of males. Maturing rapidly, females breed profusely so that the number of these insects multiplies quickly. Winged females may develop later in the season, allowing the insects to colonize new plants. In temperate regions, a phase of sexual reproduction occurs in the autumn, with the insects often overwintering as eggs.

Contents

The life cycle of some species involves an alternation between two species of host plants, for example between an annual crop and a woody plant. Some species feed on only one type of plant, while others are generalists, colonizing many plant groups. About 5,000 species of aphid have been described, all included in the family Aphididae. Around 400 of these are found on food and fiber crops, and many are serious pests of agriculture and forestry, as well as an annoyance for gardeners. So-called dairying ants have a mutualistic relationship with aphids, tending them for their honeydew and protecting them from predators.

Aphids are among the most destructive insect pests on cultivated plants in temperate regions. In addition to weakening the plant by sucking sap, they act as vectors for plant viruses and disfigure ornamental plants with deposits of honeydew and the subsequent growth of sooty moulds. Because of their ability to rapidly increase in numbers by asexual reproduction and telescopic development, they are a highly successful group of organisms from an ecological standpoint. [1]

Control of aphids is not easy. Insecticides do not always produce reliable results, given resistance to several classes of insecticide and the fact that aphids often feed on the undersides of leaves. On a garden scale, water jets and soap sprays are quite effective. Natural enemies include predatory ladybugs, hoverfly larvae, parasitic wasps, aphid midge larvae, crab spiders, lacewing larvae, and entomopathogenic fungi. An integrated pest management strategy using biological pest control can work, but is difficult to achieve except in enclosed environments such as greenhouses.

Distribution

Aphids are distributed worldwide, but are most common in temperate zones. In contrast to many taxa, aphid species diversity is much lower in the tropics than in the temperate zones. [2] They can migrate great distances, mainly through passive dispersal by winds. Winged aphids may also rise up in the day as high as 600 m where they are transported by strong winds. [3] [4] For example, the currant-lettuce aphid, Nasonovia ribisnigri , is believed to have spread from New Zealand to Tasmania around 2004 through easterly winds. [5] Aphids have also been spread by human transportation of infested plant materials, making some species nearly cosmopolitan in their distribution. [6]

Evolution

Forewing of the early Middle Triassic (early Anisian) aphid Vosegus triassicus Vosegus.jpg
Forewing of the early Middle Triassic (early Anisian) aphid Vosegus triassicus
An aphid fossilised in Baltic amber (Eocene) Baltic amber inclusions - Aphid (Hemiptera, Sternorrhyncha, Aphidoidea)7.JPG
An aphid fossilised in Baltic amber (Eocene)

Fossil history

Aphids, and the closely related adelgids and phylloxerans, probably evolved from a common ancestor some 280  million years ago, in the Early Permian period. [8] They probably fed on plants like Cordaitales or Cycadophyta. With their soft bodies, aphids do not fossilize well, and the oldest known fossil is of the species Triassoaphis cubitus from the Triassic. [9] They do however sometimes get stuck in plant exudates which solidify into amber. In 1967, when Professor Ole Heie wrote his monograph Studies on Fossil Aphids, about sixty species have been described from the Triassic, Jurassic, Cretaceous and mostly the Tertiary periods, with Baltic amber contributing another forty species. [10] The total number of species was small, but increased considerably with the appearance of the angiosperms 160  million years ago, as this allowed aphids to specialise, the speciation of aphids going hand-in-hand with the diversification of flowering plants. The earliest aphids were probably polyphagous, with monophagy developing later. [11] It has been hypothesized that the ancestors of the Adelgidae lived on conifers while those of the Aphididae fed on the sap of Podocarpaceae or Araucariaceae that survived extinctions in the late Cretaceous. Organs like the cornicles did not appear until the Cretaceous period. [8] [12] One study alternatively suggests that ancestral aphids may have lived on angiosperm bark and that feeding on leaves may be a derived trait. The Lachninae have long mouth parts that are suitable for living on bark and it has been suggested that the mid-Cretaceous ancestor fed on the bark of angiosperm trees, switching to leaves of conifer hosts in the late Cretaceous. [13] The Phylloxeridae may well be the oldest family still extant, but their fossil record is limited to the Lower Miocene Palaeophylloxera . [14]

Taxonomy

Late 20th-century reclassification within the Hemiptera reduced the old taxon "Homoptera" to two suborders: Sternorrhyncha (aphids, whiteflies, scales, psyllids, etc.) and Auchenorrhyncha (cicadas, leafhoppers, treehoppers, planthoppers, etc.) with the suborder Heteroptera containing a large group of insects known as the true bugs. The infraorder Aphidomorpha within the Sternorrhyncha varies with circumscription with several fossil groups being especially difficult to place but includes the Adelgoidea, the Aphidoidea and the Phylloxeroidea. [15] Some authors use a single superfamily Aphidoidea within which the Phylloxeridae and Adelgidae are also included while others have Aphidoidea with a sister superfamily Phylloxeroidea within which the Adelgidae and Phylloxeridae are placed. [16] Early 21st-century reclassifications substantially rearranged the families within Aphidoidea: some old families were reduced to subfamily rank (e.g., Eriosomatidae), and many old subfamilies were elevated to family rank. The most recent authoritative classifications have three superfamilies Adelgoidea, Phylloxeroidea and Aphidoidea. The Aphidoidea includes a single large family Aphididae that includes all the ~5000 [2] extant species. [17]

Phylogeny

External

Aphids, adelgids, and phylloxerids are very closely related within the suborder Sternorrhyncha, the plant-sucking bugs. They are either placed in the insect superfamily Aphidoidea [18] or into the superfamily Phylloxeroidea which contains the family Adelgidae and the family Phylloxeridae. [11] Like aphids, phylloxera feed on the roots, leaves, and shoots of grape plants, but unlike aphids, do not produce honeydew or cornicle secretions. [19] Phylloxera (Daktulosphaira vitifoliae) are insects which caused the Great French Wine Blight that devastated European viticulture in the 19th century. Similarly, adelgids or woolly conifer aphids, also feed on plant phloem and are sometimes described as aphids, but are more properly classified as aphid-like insects, because they have no cauda or cornicles. [20]

The treatment of the groups especially concerning fossil groups varies greatly due to difficulties in resolving relationships. Most modern treatments include the three superfamilies, the Adelogidea, the Aphidoidea, and the Phylloxeroidea within the infraorder Aphidomorpha along with several fossil groups. [21]

Sternorrhyncha

Psylloidea (jumping plant lice, etc) Psyllia pyricola.png

Aleyrodoidea (whiteflies) Neomaskellia bergii from CSIRO.jpg

Coccoidea (scale insects) Ceroplastes ceriferus from CSIRO.jpg

Aphidomorpha
Phylloxeroidea

Phylloxeridae (phylloxerans) Daktulosphaira vitifoliae from CSIRO.jpg

Adelgidae (woolly conifer aphids) Chermes pinifoliae.png

Aphidoidea

Aphididae (aphids) Aphid icon.png

Internal

The phylogenetic tree, based on Papasotiropoulos 2013 and Kim 2011, with additions from Ortiz-Rivas and Martinez-Torres 2009, shows the internal phylogeny of the Aphididae. [22] [23] [24]

It has been suggested that the phylogeny of the aphid groups might be revealed by examining the phylogeny of their bacterial endosymbionts, especially the obligate endosymbiont Buchnera . The results depend on the assumption that the symbionts are strictly transmitted vertically through the generations. This assumption is well supported by the evidence, and several phylogenetic relationships have been suggested on the basis of endosymbiont studies. [25] [26] [27]

Aphididae

Lachninae Schizolachnus sp 20120528.JPG

Hormaphidinae Witch Hazel Cone Gall Aphid - Flickr - treegrow (2).jpg

Calaphidinae Calaphis.flava.-.lindsey.jpg

Chaitophorinae Periphyllus sp. aphids on sycamore (cropped).jpg

Eriosomatinae (woolly aphids) Eriosoma ulmi (European Elm Leafcurl Aphid).jpg

Anoeciinae Anoecia corni1.jpg

Capitophorus , Pterocoma

Aphidinae

Macrosiphini Macrosiphum rosae (alate i.e. winged form) on a rose bud.jpg

Aphidini

Rhopalosiphina Snodgrass Rhopalosiphum prunifoliae (wings closed).jpg

Aphidina ( Aphis spp) Aphids May 2010-3.jpg


Anatomy

Front view of wheat aphid, Schizaphis graminum, showing the piercing-sucking mouthparts Schizaphis graminum usda (cropped).jpg
Front view of wheat aphid, Schizaphis graminum , showing the piercing-sucking mouthparts

Most aphids have soft bodies, which may be green, black, brown, pink, or almost colorless. Aphids have antennae with two short, broad basal segments and up to four slender terminal segments. They have a pair of compound eyes, with an ocular tubercle behind and above each eye, made up of three lenses (called triommatidia). [11] They feed on sap using sucking mouthparts called stylets, enclosed in a sheath called a rostrum, which is formed from modifications of the mandible and maxilla of the insect mouthparts. [28]

They have long, thin legs with two-jointed, two-clawed tarsi. The majority of aphids are wingless, but winged forms are produced at certain times of year in many species. Most aphids have a pair of cornicles (siphunculi), abdominal tubes on the dorsal surface of their fifth abdominal segment, through which they exude droplets of a quick-hardening defensive fluid [28] containing triacylglycerols, called cornicle wax. Other defensive compounds can also be produced by some species. [20] Aphids have a tail-like protrusion called a cauda above their rectal apertures. [11] [29] They have lost their Malpighian tubules. [30]

When host plant quality becomes poor or conditions become crowded, some aphid species produce winged offspring (alates) that can disperse to other food sources. The mouthparts or eyes can be small or missing in some species and forms. [20]

Diet

Woolly Aphids on Apple tree branch Woolly Aphids.jpg
Woolly Aphids on Apple tree branch

Many aphid species are monophagous (that is, they feed on only one plant species). Others, like the green peach aphid, feed on hundreds of plant species across many families. About 10% of species feed on different plants at different times of the year. [31]

A new host plant is chosen by a winged adult by using visual cues, followed by olfaction using the antennae; if the plant smells right, the next action is probing the surface upon landing. The stylus is inserted and saliva secreted, the sap is sampled, the xylem may be tasted and finally, the phloem is tested. Aphid saliva may inhibit phloem-sealing mechanisms and has pectinases that ease penetration. [32] Non-host plants can be rejected at any stage of the probe, but the transfer of viruses occurs early in the investigation process, at the time of the introduction of the saliva, so non-host plants can become infected. [31]

Aphids usually feed passively on sap of phloem vessels in plants, as do many of other hemipterans such as scale insects and cicadas. Once a phloem vessel is punctured, the sap, which is under pressure, is forced into the aphid's food canal. Occasionally, aphids also ingest xylem sap, which is a more dilute diet than phloem sap as the concentrations of sugars and amino acids are 1% of those in the phloem. [33] [34] Xylem sap is under negative hydrostatic pressure and requires active sucking, suggesting an important role in aphid physiology. [35] As xylem sap ingestion has been observed following a dehydration period, aphids are thought to consume xylem sap to replenish their water balance; the consumption of the dilute sap of xylem permitting aphids to rehydrate. [36] However, recent data showed aphids consume more xylem sap than expected and they notably do so when they are not dehydrated and when their fecundity decreases. This suggests aphids, and potentially, all the phloem-sap feeding species of the order Hemiptera, consume xylem sap for reasons other than replenishing water balance. [37] Although aphids passively take in phloem sap, which is under pressure, they can also draw fluid at negative or atmospheric pressure using the cibarial-pharyngeal pump mechanism present in their head. [38]

Xylem sap consumption may be related to osmoregulation. [37] High osmotic pressure in the stomach, caused by high sucrose concentration, can lead to water transfer from the hemolymph to the stomach, thus resulting in hyperosmotic stress and eventually to the death of the insect. Aphids avoid this fate by osmoregulating through several processes. Sucrose concentration is directly reduced by assimilating sucrose toward metabolism and by synthesizing oligosaccharides from several sucrose molecules, thus reducing the solute concentration and consequently the osmotic pressure. [39] [40] Oligosaccharides are then excreted through honeydew, explaining its high sugar concentrations, which can then be used by other animals such as ants. Furthermore, water is transferred from the hindgut, where osmotic pressure has already been reduced, to the stomach to dilute stomach content. [41] Eventually, aphids consume xylem sap to dilute the stomach osmotic pressure. [37] All these processes function synergetically, and enable aphids to feed on high-sucrose-concentration plant sap, as well as to adapt to varying sucrose concentrations. [37]

Plant sap is an unbalanced diet for aphids, as it lacks essential amino acids, which aphids, like all animals, cannot synthesise, and possesses a high osmotic pressure due to its high sucrose concentration. [34] [42] Essential amino acids are provided to aphids by bacterial endosymbionts, harboured in special cells, bacteriocytes. [43] These symbionts recycle glutamate, a metabolic waste of their host, into essential amino acids. [44] [45]

Carotenoids and photoheterotrophy

Some species of aphids have acquired the ability to synthesise red carotenoids by horizontal gene transfer from fungi. [46] They are the only animals other than two-spotted spider mites and the oriental hornet with this capability. [47] Using their carotenoids, aphids may well be able to absorb solar energy and convert it to a form that their cells can use, ATP. This is the only known example of photoheterotrophy in animals. The carotene pigments in aphids form a layer close to the surface of the cuticle, ideally placed to absorb sunlight. The excited carotenoids seem to reduce NAD to NADH which is oxidized in the mitochondria for energy. [48]

Reproduction

Soybean aphid alternates between hosts and between asexual and sexual reproduction. Soybeanaphidlifecycle.gif
Soybean aphid alternates between hosts and between asexual and sexual reproduction.

The simplest reproductive strategy is for an aphid to have a single host all year round. On this it may alternate between sexual and asexual generations (holocyclic) or alternatively, all young may be produced by parthenogenesis, eggs never being laid (anholocyclic). Some species can have both holocyclic and anholocyclic populations under different circumstances but no known aphid species reproduce solely by sexual means. [50] The alternation of sexual and asexual generations may have evolved repeatedly. [51]

However, aphid reproduction is often more complex than this and involves migration between different host plants. In about 10% of species, there is an alternation between woody (primary hosts) on which the aphids overwinter and herbaceous (secondary) host plants, where they reproduce abundantly in the summer. [20] [50] A few species can produce a soldier caste, other species show extensive polyphenism under different environmental conditions and some can control the sex ratio of their offspring depending on external factors. [52]

When a typical sophisticated reproductive strategy is used, only females are present in the population at the beginning of the seasonal cycle (although a few species of aphids have been found to have both male and female sexes at this time). The overwintering eggs that hatch in the spring result in females, called fundatrices (stem mothers). Reproduction typically does not involve males (parthenogenesis) and results in a live birth (viviparity). [53] The live young are produced by pseudoplacental viviparity, which is the development of eggs, deficient in the yolk, the embryos fed by a tissue acting as a placenta. The young emerge from the mother soon after hatching. [54]

Eggs are parthenogenetically produced without meiosis [55] [53] and the offspring are clonal to their mother, so they are all female (thelytoky). [11] [54] The embryos develop within the mothers' ovarioles, which then give birth to live (already hatched) first-instar female nymphs. As the eggs begin to develop immediately after ovulation, an adult female can house developing female nymphs which already have parthenogenetically developing embryos inside them (i.e. they are born pregnant). This telescoping of generations enables aphids to increase in number with great rapidity. The offspring resemble their parent in every way except size. Thus, a female's diet can affect the body size and birth rate of more than two generations (daughters and granddaughters). [11] [56] [57]

This process repeats itself throughout the summer, producing multiple generations that typically live 20 to 40 days. For example, some species of cabbage aphids (like Brevicoryne brassicae ) can produce up to 41 generations of females in a season. Thus, one female hatched in spring can theoretically produce billions of descendants, were they all to survive. [58]

Aphid giving birth to live young: populations are often entirely female. Aphid-giving-birth.jpg
Aphid giving birth to live young: populations are often entirely female.

In autumn, aphids reproduce sexually and lay eggs. Environmental factors such as a change in photoperiod and temperature, or perhaps a lower food quantity or quality, causes females to parthenogenetically produce sexual females and males. [55] The males are genetically identical to their mothers except that, with the aphids' X0 sex-determination system, they have one fewer sex chromosome. [55] These sexual aphids may lack wings or even mouthparts. [20] Sexual females and males mate, and females lay eggs that develop outside the mother. The eggs survive the winter and hatch into winged (alate) or wingless females the following spring. This occurs in, for example, the life cycle of the rose aphid (Macrosiphum rosae), which may be considered typical of the family. However, in warm environments, such as in the tropics or a greenhouse, aphids may go on reproducing asexually for many years. [28]

Aphids reproducing asexually by parthenogenesis can have genetically identical winged and non-winged female progeny. Control is complex; some aphids alternate during their life-cycles between genetic control (polymorphism) and environmental control (polyphenism) of production of winged or wingless forms. [59] Winged progeny tend to be produced more abundantly under unfavorable or stressful conditions. Some species produce winged progeny in response to low food quality or quantity. e.g. when a host plant is starting to senesce. [60]  The winged females migrate to start new colonies on a new host plant. For example, the apple aphid (Aphis pomi), after producing many generations of wingless females gives rise to winged forms that fly to other branches or trees of its typical food plant. [61] Aphids that are attacked by ladybugs, lacewings, parasitoid wasps, or other predators can change the dynamics of their progeny production. When aphids are attacked by these predators, alarm pheromones, in particular beta-farnesene, are released from the cornicles. These alarm pheromones cause several behavioral modifications that, depending on the aphid species, can include walking away and dropping off the host plant. Additionally, alarm pheromone perception can induce the aphids to produce winged progeny that can leave the host plant in search of a safer feeding site. [62] Viral infections, which can be extremely harmful to aphids, can also lead to the production of winged offspring. [63] For example, Densovirus infection has a negative impact on rosy apple aphid ( Dysaphis plantaginea ) reproduction, but contributes to the development of aphids with wings, which can transmit the virus more easily to new host plants. [64] Additionally, symbiotic bacteria that live inside of the aphids can also alter aphid reproductive strategies based on the exposure to environmental stressors. [65]

The life stages of the green apple aphid (Aphis pomi). Drawing by Robert Evans Snodgrass, 1930 Snodgrass Aphis pomi.jpg
The life stages of the green apple aphid (Aphis pomi). Drawing by Robert Evans Snodgrass, 1930

In the autumn, host-alternating (heteroecious) aphid species produce a special winged generation that flies to different host plants for the sexual part of the life cycle. Flightless female and male sexual forms are produced and lay eggs. [66] Some species such as Aphis fabae (black bean aphid), Metopolophium dirhodum (rose-grain aphid), Myzus persicae (peach-potato aphid), and Rhopalosiphum padi (bird cherry-oat aphid) are serious pests. They overwinter on primary hosts on trees or bushes; in summer, they migrate to their secondary host on a herbaceous plant, often a crop, then the gynoparae return to the tree in autumn. Another example is the soybean aphid (Aphis glycines). As fall approaches, the soybean plants begin to senesce from the bottom upwards. The aphids are forced upwards and start to produce winged forms, first females and later males, which fly off to the primary host, buckthorn. Here they mate and overwinter as eggs. [49]

Ecology

Ant mutualism

An ant guards its aphids Ant guards its Aphids.jpg
An ant guards its aphids
Ants tending aphids Myrmica puceron.jpg
Ants tending aphids

Some species of ants farm aphids, protecting them on the plants where they are feeding, and consuming the honeydew the aphids release from the terminations of their alimentary canals. This is a mutualistic relationship, with these dairying ants milking the aphids by stroking them with their antennae. [lower-alpha 2] [67] Although mutualistic, the feeding behaviour of aphids is altered by ant attendance. Aphids attended by ants tend to increase the production of honeydew in smaller drops with a greater concentration of amino acids. [68]

Some farming ant species gather and store the aphid eggs in their nests over the winter. In the spring, the ants carry the newly hatched aphids back to the plants. Some species of dairying ants (such as the European yellow meadow ant, Lasius flavus) [69] manage large herds of aphids that feed on roots of plants in the ant colony. Queens leaving to start a new colony take an aphid egg to found a new herd of underground aphids in the new colony. These farming ants protect the aphids by fighting off aphid predators. [67] Some bees in coniferous forests collect aphid honeydew to make forest honey. [28]

Ant extracting honeydew from an aphid Ant feeding on honeydew.JPG
Ant extracting honeydew from an aphid

An interesting variation in ant–aphid relationships involves lycaenid butterflies and Myrmica ants. For example, Niphanda fusca butterflies lay eggs on plants where ants tend herds of aphids. The eggs hatch as caterpillars which feed on the aphids. The ants do not defend the aphids from the caterpillars, since the caterpillars produce a pheromone which deceives the ants into treating them like ants, and carrying the caterpillars into their nest. Once there, the ants feed the caterpillars, which in return produce honeydew for the ants. When the caterpillars reach full size, they crawl to the colony entrance and form cocoons. After two weeks, the adult butterflies emerge and take flight. At this point, the ants attack the butterflies, but the butterflies have a sticky wool-like substance on their wings that disables the ants' jaws, allowing the butterflies to fly away without being harmed. [70]

Another ant-mimicking gall aphid, Paracletus cimiciformis (Eriosomatinae), has evolved a complex double strategy involving two morphs of the same clone and Tetramorium ants. Aphids of the round morph cause the ants to farm them, as with many other aphids. The flat morph aphids are aggressive mimics with a "wolf in sheep's clothing" strategy: they have hydrocarbons in their cuticle that mimic those of the ants, and the ants carry them into the brood chamber of the ants' nest and raise them like ant larvae. Once there, the flat morph aphids behave like predators, drinking the body fluids of ant larvae. [71]

Bacterial endosymbiosis

Endosymbiosis with micro-organisms is common in insects, with more than 10% of insect species relying on intracellular bacteria for their development and survival. [72] Aphids harbour a vertically transmitted (from parent to its offspring) obligate symbiosis with Buchnera aphidicola , the primary symbiont, inside specialized cells, the bacteriocytes. [73] Five of the bacteria genes have been transferred to the aphid nucleus. [74] The original association is estimated to have occurred in a common ancestor 280 to 160 million years ago and enabled aphids to exploit a new ecological niche, feeding on phloem-sap of vascular plants. B. aphidicola provides its host with essential amino acids, which are present in low concentrations in plant sap. [75] The metabolites from endosymbionts are also excreted in honeydew. [76] The stable intracellular conditions, as well as the bottleneck effect experienced during the transmission of a few bacteria from the mother to each nymph, increase the probability of transmission of mutations and gene deletions. [77] [78] As a result, the size of the B. aphidicola genome is greatly reduced, compared to its putative ancestor. [79] Despite the apparent loss of transcription factors in the reduced genome, gene expression is highly regulated, as shown by the ten-fold variation in expression levels between different genes under normal conditions. [80] Buchnera aphidicola gene transcription, although not well understood, is thought to be regulated by a small number of global transcriptional regulators and/or through nutrient supplies from the aphid host. [81]

Some aphid colonies also harbour secondary or facultative (optional extra) bacterial symbionts. These are vertically transmitted, and sometimes also horizontally (from one lineage to another and possibly from one species to another). [82] [83] So far, the role of only some of the secondary symbionts has been described; Regiella insecticola plays a role in defining the host-plant range, [84] [85] Hamiltonella defensa provides resistance to parasitoids but only when it is in turn infected by the bacteriophage APSE, [86] [87] and Serratia symbiotica prevents the deleterious effects of heat. [88]

Predators

Aphids are eaten by many bird and insect predators. In a study on a farm in North Carolina, six species of passerine bird consumed nearly a million aphids per day between them, the top predators being the American goldfinch, with aphids forming 83% of its diet, and the vesper sparrow. [89] Insects that attack aphids include the adults and larvae of predatory ladybirds, hoverfly larvae, parasitic wasps, aphid midge larvae, "aphid lions" (the larvae of green lacewings), and arachnids such as spiders. Among ladybirds, Myzia oblongoguttata is a dietary specialist which only feeds on conifer aphids, whereas Adalia bipunctata and Coccinella septempunctata are generalists, feeding on large numbers of species. The eggs are laid in batches, each female laying several hundred. Female hoverflies lay several thousand eggs. The adults feed on pollen and nectar but the larvae feed voraciously on aphids; Eupeodes corollae adjusts the number of eggs laid to the size of the aphid colony. [90]

Predators eating aphids

Aphids are often infected by bacteria, viruses, and fungi. They are affected by the weather, such as precipitation, [91] temperature [92] and wind. [93] Fungi that attack aphids include Neozygites fresenii , Entomophthora , Beauveria bassiana , Metarhizium anisopliae , and entomopathogenic fungi such as Lecanicillium lecanii . Aphids brush against the microscopic spores. These stick to the aphid, germinate, and penetrate the aphid's skin. The fungus grows in the aphid's hemolymph. After about three days, the aphid dies and the fungus releases more spores into the air. Infected aphids are covered with a woolly mass that progressively grows thicker until the aphid is obscured. Often, the visible fungus is not the one that killed the aphid, but a secondary infection. [91]

Aphids can be easily killed by unfavourable weather, such as late spring freezes. [94] Excessive heat kills the symbiotic bacteria that some aphids depend on, which makes the aphids infertile. [95] Rain prevents winged aphids from dispersing, and knocks aphids off plants and thus kills them from the impact or by starvation, [91] [96] [97] but cannot be relied on for aphid control. [98]

Anti-predator defences

Aphid secreting defensive fluid from the cornicles Aphid-sap2.jpg
Aphid secreting defensive fluid from the cornicles

Most aphids have little protection from predators. Some species interact with plant tissues forming a gall, an abnormal swelling of plant tissue. Aphids can live inside the gall, which provides protection from predators and the elements. A number of galling aphid species are known to produce specialised "soldier" forms, sterile nymphs with defensive features which defend the gall from invasion. [28] [99] [100] For example, Alexander's horned aphids are a type of soldier aphid that has a hard exoskeleton and pincer-like mouthparts. [70] :144 A woolly aphid, Colophina clematis , has first instar "soldier" nymphs that protect the aphid colony, killing larvae of ladybirds, hoverflies and the flower bug Anthocoris nemoralis by climbing on them and inserting their stylets. [101]

Although aphids cannot fly for most of their life cycle, they can escape predators and accidental ingestion by herbivores by dropping off the plant onto the ground. [102] Others species use the soil as a permanent protection, feeding on the vascular systems of roots and remaining underground all their lives. They are often attended by ants, for the honeydew they produce and are carried from plant to plant by the ants through their tunnels. [89]

Some species of aphid, known as "woolly aphids" (Eriosomatinae), excrete a "fluffy wax coating" for protection. [28] The cabbage aphid, Brevicoryne brassicae , sequesters secondary metabolites from its host, stores them and releases chemicals that produce a violent chemical reaction and strong mustard oil smell to repel predators. [103] Peptides produced by aphids, Thaumatins, are thought to provide them with resistance to some fungi. [104]

It was common at one time to suggest that the cornicles were the source of the honeydew, and this was even included in the Shorter Oxford English Dictionary [105] and the 2008 edition of the World Book Encyclopedia . [106] In fact, honeydew secretions are produced from the anus of the aphid, [107] while cornicles mostly produce defensive chemicals such as waxes. There also is evidence of cornicle wax attracting aphid predators in some cases. [108]

Some clones of Aphis craccivora are sufficiently toxic to the invasive and dominant predatory ladybird Harmonia axyridis to suppress it locally, favouring other ladybird species; the toxicity is in this case narrowly specific to the dominant predator species. [109]

Parasitoids

Aphids are abundant and widespread, and serve as hosts to a large number of parasitoids, many of them being very small (c. 0.1 inches (2.5 mm) long) parasitoid wasps. [110] One species, Aphis ruborum , for example, is host to at least 12 species of parasitoid wasps. [111] Parasitoids have been investigated intensively as biological control agents, and many are used commercially for this purpose. [112]

Plant-aphid interactions

Aphids on plant host 2012.10.19.-25-Mannheim Vogelstang-Blattlaeuse.jpg
Aphids on plant host

Plants mount local and systemic defenses against aphid attack. Young leaves in some plants contain chemicals that discourage attack while the older leaves have lost this resistance, while in other plant species, resistance is acquired by older tissues and the young shoots are vulnerable. Volatile products from interplanted onions have been shown to prevent aphid attack on adjacent potato plants by encouraging the production of terpenoids, a benefit exploited in the traditional practice of companion planting, while plants neighboring infested plants showed increased root growth at the expense of the extension of aerial parts. [31] The wild potato, Solanum berthaultii , produces an aphid alarm pheromone, (E)-β-farnesene, as an allomone, a pheromone to ward off attack; it effectively repels the aphid Myzus persicae at a range of up to 3 millimetres. [113] S. berthaultii and other wild potato species have a further anti-aphid defence in the form of glandular hairs which, when broken by aphids, discharge a sticky liquid that can immobilise some 30% of the aphids infesting a plant. [114]

Plants exhibiting aphid damage can have a variety of symptoms, such as decreased growth rates, mottled leaves, yellowing, stunted growth, curled leaves, browning, wilting, low yields, and death. The removal of sap creates a lack of vigor in the plant, and aphid saliva is toxic to plants. Aphids frequently transmit plant viruses to their hosts, such as to potatoes, cereals, sugarbeets, and citrus plants. [28] The green peach aphid, Myzus persicae , is a vector for more than 110 plant viruses. Cotton aphids (Aphis gossypii) often infect sugarcane, papaya and peanuts with viruses. [20] In plants which produce the phytoestrogen coumestrol, such as alfalfa, damage by aphids is linked with higher concentrations of coumestrol. [115]

Aphid with honeydew, from the anus, not the cornicles Blattlaus mit Honigtau 6644.jpg
Aphid with honeydew, from the anus, not the cornicles

The coating of plants with honeydew can contribute to the spread of fungi which can damage plants. [116] [117] Honeydew produced by aphids has been observed to reduce the effectiveness of fungicides as well. [118]

A hypothesis that insect feeding may improve plant fitness was floated in the mid-1970s by Owen and Wiegert. It was felt that the excess honeydew would nourish soil micro-organisms, including nitrogen fixers. In a nitrogen-poor environment, this could provide an advantage to an infested plant over an uninfested plant. However, this does not appear to be supported by observational evidence. [119]

Sociality

Some aphids show some of the traits of eusociality, joining insects such as ants, bees, and termites. However, there are differences between these sexual social insects and the clonal aphids, which are all descended from a single female parthenogenetically and share an identical genome. About fifty species of aphid, scattered among the closely related, host-alternating lineages Eriosomatinae and Hormaphidinae, have some type of defensive morph. These are gall-creating species, with the colony living and feeding inside a gall that they form in the host's tissues. Among the clonal population of these aphids, there may be several distinct morphs and this lays the foundation for a possible specialization of function, in this case, a defensive caste. The soldier morphs are mostly first and second instars with the third instar being involved in Eriosoma moriokense and only in Smythurodes betae are adult soldiers known. The hind legs of soldiers are clawed, heavily sclerotized and the stylets are robust making it possible to rupture and crush small predators. [120] The larval soldiers are altruistic individuals, unable to advance to breeding adults but acting permanently in the interests of the colony. Another requirement for the development of sociality is provided by the gall, a colonial home to be defended by the soldiers. [121]

The soldiers of gall-forming aphids also carry out the job of cleaning the gall. The honeydew secreted by the aphids is coated in a powdery wax to form "liquid marbles" [122] that the soldiers roll out of the gall through small orifices. [100] Aphids that form closed galls use the plant's vascular system for their plumbing: the inner surfaces of the galls are highly absorbent and wastes are absorbed and carried away by the plant. [100]

Interactions with humans

Pest status

About 5000 species of aphid have been described and of these, some 450 species have colonized food and fiber crops. As direct feeders on plant sap, they damage crops and reduce yields, but they have a greater impact by being vectors of plant viruses. The transmission of these viruses depends on the movements of aphids between different parts of a plant, between nearby plants, and further afield. In this respect, the probing behavior of an aphid tasting a host is more damaging than lengthy aphid feeding and reproduction by stay-put individuals. The movement of aphids influences the timing of virus epidemics. [123] They are major pests of greenhouse crops and species often encountered in greenhouses include: green peach aphid ( Myzus persicae ), cotton or melon aphid ( Aphis gossypii ), potato aphid ( Macrosiphum euphorbiae ), foxglove aphid ( Aulacorthum solani ) and chrysanthemum aphid ( Macrosiphoniella sanborni ) and others, which cause leaf yellowing, distorted leaves, and plant stunting; the excreted honeydew is a growing medium for a number of fungal pathogens including black sooty molds from the genera Capnodium , Fumago , and Scorias which then infect leaves and inhibit growth by reducing photosynthesis. [124]

Aphids, especially during large outbreaks, have been known to trigger allergic inhalant reactions in sensitive humans. [125]

Dispersal can be by walking or flight, appetitive dispersal, or by migration. Winged aphids are weak fliers, lose their wings after a few days and only fly by day. Dispersal by flight is affected by the impact, air currents, gravity, precipitation, and other factors, or dispersal may be accidental, caused by the movement of plant materials, animals, farm machinery, vehicles, or aircraft. [123]

Control

Parasitoid braconid wasp ovipositing in black bean aphid Wasp & aphid May 2010-1.jpg
Parasitoid braconid wasp ovipositing in black bean aphid

Insecticide control of aphids is difficult, as they breed rapidly, so even small areas missed may enable the population to recover promptly. Aphids may occupy the undersides of leaves where spray misses them, while systemic insecticides do not move satisfactorily into flower petals. Finally, some aphid species are resistant to common insecticide classes including carbamates, organophosphates, and pyrethroids. [126]

For small backyard infestations, spraying plants thoroughly with a strong water jet every few days may be sufficient protection. An insecticidal soap solution can be an effective household remedy to control aphids, but it only kills aphids on contact and has no residual effect. Soap spray may damage plants, especially at higher concentrations or at temperatures above 32 °C (90 °F); some plant species are sensitive to soap sprays. [112] [127] [128]

Green peach aphid, Myzus persicae, killed by the fungus Pandora neoaphidis (Entomophthorales) Pandora neoaphidis.jpg
Green peach aphid, Myzus persicae , killed by the fungus Pandora neoaphidis (Entomophthorales)

Aphid populations can be sampled using yellow-pan or Moericke traps. These are yellow containers with water that attract aphids. [129] Aphids respond positively to green and their attraction to yellow may not be a true colour preference but related to brightness. Their visual receptors peak in sensitivity from 440 to 480 nm and are insensitive in the red region. Moericke found that aphids avoided landing on white coverings placed on soil and were repelled even more by shiny aluminium surfaces. [130] Integrated pest management of various species of aphids can be achieved using biological insecticides based on fungi such as Lecanicillium lecanii, Beauveria bassiana or Isaria fumosorosea . [131] Fungi are the main pathogens of aphids; Entomophthorales can quickly cut aphid numbers in nature. [132]

Aphids may also be controlled by the release of natural enemies, in particular lady beetles and parasitoid wasps. However, since adult lady beetles tend to fly away within 48 hours after release, without laying eggs, repeated applications of large numbers of lady beetles are needed to be effective. For example, one large, heavily infested rose bush may take two applications of 1500 beetles each. [112] [133]

The ability to produce allomones such as farnesene to repel and disperse aphids and to attract their predators has been experimentally transferred to transgenic Arabidopsis thaliana plants using an Eβf synthase gene in the hope that the approach could protect transgenic crops. [134] Eβ farnesene has however found to be ineffective in crop situations although stabler synthetic forms help improve the effectiveness of control using fungal spores and insecticides through increased uptake caused by movements of aphids. [135]

In human culture

Aphids are familiar to farmers and gardeners, mainly as pests. Peter Marren and Richard Mabey record that Gilbert White described an invading "army" of black aphids that arrived in his village of Selborne, Hampshire, England, in August 1774 in "great clouds", covering every plant, while in the unusually hot summer of 1783, White found that honeydew was so abundant as to "deface and destroy the beauties of my garden", though he thought the aphids were consuming rather than producing it. [136]

Infestation of the Chinese sumac ( Rhus chinensis ) by Chinese sumac aphids ( Schlechtendalia chinensis ) can create "Chinese galls" which are valued as a commercial product. As "Galla Chinensis", they are used in traditional Chinese medicine to treat coughs, diarrhea, night sweats, dysentery and to stop intestinal and uterine bleeding. Chinese galls are also an important source of tannins. [28]

See also

Notes

  1. The term "black fly" is also used for the Simuliidae, among them the vector of river blindness.
  2. Dairying ants also milk mealybugs and other insects.

Related Research Articles

<span class="mw-page-title-main">Hemiptera</span> Order of insects often called true bugs

Hemiptera is an order of insects, commonly called true bugs, comprising over 80,000 species within groups such as the cicadas, aphids, planthoppers, leafhoppers, assassin bugs, bed bugs, and shield bugs. They range in size from 1 mm (0.04 in) to around 15 cm (6 in), and share a common arrangement of piercing-sucking mouthparts. The name "true bugs" is often limited to the suborder Heteroptera.

<span class="mw-page-title-main">Aphididae</span> Family of true bugs

The Aphididae are a very large insect family in the aphid superfamily (Aphidoidea), of the order Hemiptera. These insects suck the sap from plant leaves. Several thousand species are placed in this family, many of which are considered plant/crop pests. They are the family of insects containing most plant virus vectors with the green peach aphid being one of the most prevalent and indiscriminate carriers.

<span class="mw-page-title-main">Eriosomatinae</span> Subfamily of aphids

Woolly aphids are sap-sucking insects that produce a filamentous waxy white covering which resembles cotton or wool. The adults are winged and move to new locations where they lay egg masses. The nymphs often form large cottony masses on twigs, for protection from predators.

<i>Buchnera aphidicola</i> Species of bacterium

Buchnera aphidicola, a member of the Pseudomonadota and the only species in the genus Buchnera, is the primary endosymbiont of aphids, and has been studied in the pea aphid, Acyrthosiphon pisum. Buchnera is believed to have had a free-living, Gram-negative ancestor similar to a modern Enterobacterales, such as Escherichia coli. Buchnera is 3 µm in diameter and has some of the key characteristics of its Enterobacterales relatives, such as a Gram-negative cell wall. However, unlike most other Gram-negative bacteria, Buchnera lacks the genes to produce lipopolysaccharides for its outer membrane. The long association with aphids and the limitation of crossover events due to strictly vertical transmission has seen the deletion of genes required for anaerobic respiration, the synthesis of amino sugars, fatty acids, phospholipids, and complex carbohydrates. This has resulted not only in one of the smallest known genomes of any living organism, but also one of the most genetically stable.

<span class="mw-page-title-main">Honeydew (secretion)</span> Sugar-rich liquid

Honeydew is a sugar-rich sticky liquid, secreted by aphids, some scale insects, and many other true bugs and some other insects as they feed on plant sap. When their mouthpart penetrates the phloem, the sugary, high-pressure liquid is forced out of the anus of the aphid. Honeydew is particularly common as a secretion in hemipteran insects and is often the basis for trophobiosis. Some caterpillars of Lycaenidae butterflies and some moths also produce honeydew.

<span class="mw-page-title-main">Russian wheat aphid</span> Species of true bug

The Russian wheat aphid is an aphid that can cause significant losses in cereal crops. The species was introduced to the United States in 1986 and is considered an invasive species there. This aphid is pale green and up to 2 mm long. Cornicles are very short, rounded, and appear to be lacking. There is an appendage above the cauda giving the aphid the appearance of having two tails. The saliva of this aphid is toxic to the plant and causes whitish striping on cereal leaves. Feeding by this aphid will also cause the flag leaf to turn white and curl around the head causing incomplete head emergence. Its host plants are cereal grain crops including wheat and barley and to a lesser extent, wild grasses such as wheatgrasses, brome-grasses, ryegrasses and anything in the grass family.

<i>Myzus persicae</i> Aphid of peach, potato, other crops

Myzus persicae, known as the green peach aphid, greenfly, or the peach-potato aphid, is a small green aphid belonging to the order Hemiptera. It is the most significant aphid pest of peach trees, causing decreased growth, shrivelling of the leaves and the death of various tissues. It also acts as a vector for the transport of plant viruses such as cucumber mosaic virus (CMV), potato virus Y (PVY) and tobacco etch virus (TEV). Potato virus Y and potato leafroll virus can be passed to members of the nightshade/potato family (Solanaceae), and various mosaic viruses to many other food crops.

<i>Aphis gossypii</i> Species of insect

Aphis gossypii is a tiny insect, an aphid ("greenfly") in the superfamily Aphidoidea in the order Hemiptera. It is a true bug and sucks sap from plants. It is a widely distributed pest of a variety of agricultural crops in the families Cucurbitaceae, Rutaceae and Malvaceae. Common names include cotton aphid, melon aphid and melon and cotton aphid.

<span class="mw-page-title-main">Black bean aphid</span> Species of true bug

The black bean aphid is a small black insect in the genus Aphis, with a broad, soft body, a member of the order Hemiptera. Other common names include blackfly, bean aphid, and beet leaf aphid. In the warmer months of the year, it is found in large numbers on the undersides of leaves and on the growing tips of host plants, including various agricultural crops and many wild and ornamental plants. Both winged and wingless forms exist, and at this time of year, they are all females. They suck sap from stems and leaves and cause distortion of the shoots, stunted plants, reduced yield, and spoiled crops. This aphid also acts as a vector for viruses that cause plant disease, and the honeydew it secretes may encourage the growth of sooty mould. It breeds profusely by live birth, but its numbers are kept in check, especially in the later part of the summer, by various predatory and parasitic insects. Ants feed on the honeydew it produces, and take active steps to remove predators. It is a widely distributed pest of agricultural crops and can be controlled by chemical or biological means. In the autumn, winged forms move to different host plants, where both males and females are produced. These mate and the females lay eggs which overwinter.

<i>Acyrthosiphon pisum</i> Species of true bug

Acyrthosiphon pisum, commonly known as the pea aphid, is a sap-sucking insect in the family Aphididae. It feeds on several species of legumes worldwide, including forage crops, such as pea, clover, alfalfa, and broad bean, and ranks among the aphid species of major agronomical importance. The pea aphid is a model organism for biological study whose genome has been sequenced and annotated.

<i>Rhopalosiphum maidis</i> Species of true bug

Rhopalosiphum maidis, common names corn leaf aphid and corn aphid, is an insect, and a pest of maize and other crops. It has a nearly worldwide distribution and is typically found in agricultural fields, grasslands, and forest-grassland zones. Among aphids that feed on maize, it is the most commonly encountered and most economically damaging, particularly in tropical and warmer temperate areas. In addition to maize, R. maidis damages rice, sorghum, and other cultivated and wild monocots.

<i>Pemphigus spyrothecae</i> Species of true bug

Pemphigus spyrothecae, or the poplar spiral gall aphid, is a social insect which exhibits apparent altruistic behaviors. The aphids form galls and act as colony defenders, at times sacrificing their own lives to do so. It has been shown that colony defense is more likely in habitats that are difficult to obtain and can hold a large number of individuals. These gall locations are crucial because plants have a short window in which a gall can be produced. Thus, it is important for there to be a defense system that enables the aphids to retain their galls. The need for defense arises when the gall is opened up to allow winged aphid migrants to leave and to release any waste. The process of repairing the holes can take up to 10 days; during this time, the gall is susceptible to intruding predators.

<span class="mw-page-title-main">Phylloxeridae</span> Family of true bugs

Phylloxeridae is a small family of plant-parasitic hemipterans closely related to aphids with only 75 described species. This group comprises two subfamilies and 11 genera with one that is fossil. The genus type is Phylloxera. The Phylloxeridae species are usually called phylloxerans or phylloxerids.

<i>Aphis craccivora</i> Species of true bug

Aphis craccivora, variously known as the cowpea aphid, groundnut aphid or black legume aphid, is a true bug in the family Aphididae. Originally of probable Palearctic origin, it is now an invasive species of cosmopolitan distribution.

<i>Aphis pomi</i> Species of true bug

Aphis pomi, commonly known as the apple aphid, or the green apple aphid, is a true bug in the family Aphididae. It is found on young growth of apple trees and on other members of the rose family where it feeds by sucking sap. Reproduction is mainly by parthenogenesis, in which unmated females give birth to live young.

<i>Hyalopterus pruni</i> Species of true bug

Hyalopterus pruni, the mealy plum aphid, is an aphid in the superfamily Aphidoidea in the order Hemiptera. It is a true bug and sucks sap from plants.

<i>Rhopalosiphum rufiabdominale</i> Species of aphid

Rhopalosiphum rufiabdominale, the rice root aphid or red rice root aphid, is a sap-sucking insect pest with a wide host range and a global distribution. As a member of the superfamily Aphidoidea, it is one of 16 species of the genus Rhopalosiphum. Adults and nymphs are soft-bodied and usually dark green with brown, red, or yellow tones. Like all aphids, reproduction is sexual and asexual, depending on the environmental conditions and host plant. Rice root aphids cause injury to external plant parts, namely the roots or stem, by feeding on plant sap and vector several important plant viruses. The hosts of this pest extend across multiple plant families with most belonging to Rosaceae, Poaceae, and Solanaceae. R. rufiabdominale is universally associated with Prunus species but also infests various field crops, greenhouse vegetables, cannabis, and other ornamental plants. While this aphid originates from east Asia, it spans nearly every continent. Dispersal is particularly widespread across the United States, India, and Australia, with crop damage documented in multiple instances, although economic losses are primarily associated with Japanese rice crops. Nonetheless, it remains a pest of serious concern due to its high mobility, discrete habitat, and adaptive plasticity, giving it the rightful reputation as a successful invader.

<i>Eriosoma lanigerum</i> Species of true bug

Eriosoma lanigerum, the woolly apple aphid, woolly aphid or American blight, is an aphid in the superfamily Aphidoidea in the order Hemiptera. It is a true bug and sucks sap from plants.

<i>Tetraneura akinire</i> Species of true bug

The oriental grass root aphid, Tetraneura akinire, is an aphid in the superfamily Aphidoidea in the order Hemiptera. It is a true bug and sucks sap from plants.

Paracletus cimiciformis is a species of aphid with a complex life cycle. Its primary host plant is Pistacia and its secondary host is a grass, where it is present on the roots. Here it is associated with an ant and part of its life cycle is spent in the ant's nest.

References

  1. Piper, Ross (2007). Extraordinary Animals: An Encyclopedia of Curious and Unusual Animals . Greenwood Press. pp.  6–9. ISBN   978-0-313-33922-6.
  2. 1 2 Żyła, Dagmara; Homan, Agnieszka; Wegierek, Piotr (2017). "Polyphyly of the extinct family Oviparosiphidae and its implications for inferring aphid evolution (Hemiptera, Sternorrhyncha)". PLOS ONE. 12 (4): e0174791. Bibcode:2017PLoSO..1274791Z. doi: 10.1371/journal.pone.0174791 . PMC   5405925 . PMID   28445493.
  3. Berry, R. E.; Taylor, L. R. (1968). "High-Altitude Migration of Aphids in Maritime and Continental Climates". Journal of Animal Ecology. 37 (3): 713–722. doi:10.2307/3084. JSTOR   3084.
  4. Isard, Scott A.; Irwin, Michael E.; Hollinger, Steven E. (1990-10-01). "Vertical Distribution of Aphids (Homoptera: Aphididae) in the Planetary Boundary Layer". Environmental Entomology. 19 (5): 1473–1484. doi:10.1093/ee/19.5.1473.
  5. Hill, L. (2012). "The currant lettuce aphid, Nasonovia ribisnigri arrives in Tasmania: Part 1". Victorian Entomologist. 42 (2): 29–31.
  6. Margaritopoulos, John T.; Kasprowicz, Louise; Malloch, Gaynor L.; Fenton, Brian (2009-05-11). "Tracking the global dispersal of a cosmopolitan insect pest, the peach potato aphid". BMC Ecology. 9: 13. doi: 10.1186/1472-6785-9-13 . PMC   2687420 . PMID   19432979.
  7. Szwedo, J.; Nel, A. (2011). "The oldest aphid insect from the Middle Triassic of the Vosges, France". Acta Palaeontologica Polonica. 56 (4): 757–766. doi: 10.4202/app.2010.0034 .
  8. 1 2 Capinera, John L. (2008). Encyclopedia of Entomology. Springer Science & Business Media. pp. 193–194. ISBN   978-1-4020-6242-1. Archived from the original on 2020-08-05. Retrieved 2018-02-04.
  9. Johnson, Christine; Agosti, Donat; Delabie, Jocques H.; Dumpert, Klaus; Williams, D. J.; von Tschirnhaus, Michael; Macshwitz, Ulrich (2001). "Acropyga and Azteca ants (Hymenoptera: Formicidae) with scale insects (Sternorrhyncha: Coccoidea): 20 million years of intimate symbiosis" (PDF). American Museum Novitates (3335): 1–18. doi:10.1206/0003-0082(2001)335<0001:AAAAHF>2.0.CO;2. S2CID   55067700. Archived (PDF) from the original on 2012-09-27. Retrieved 2010-10-18.
  10. Russell, Louise M. (1968). "Studies on Fossil Aphids" (PDF). Bulletin of the Entomological Society of America. 14 (2): 139–140. doi:10.1093/besa/14.2.139a . Retrieved 2018-02-04.
  11. 1 2 3 4 5 6 Dixon, A. F. G. (1998). Aphid Ecology (2nd ed.). Chapman and Hall. ISBN   978-0-412-74180-7. Archived from the original on 2013-02-18. Retrieved 2016-05-24.
  12. Von Dohlen, Carol D.; Moran, Nancy A. (2000). "Molecular data support a rapid radiation of aphids in the Cretaceous and multiple origins of host alternation". Biological Journal of the Linnean Society. 71 (4): 689–717. doi: 10.1111/j.1095-8312.2000.tb01286.x .
  13. Chen, Rui; Favret, Colin; Jiang, Liyun; Wang, Zhe; Qiao, Gexia (29 September 2015). "An aphid lineage maintains a bark-feeding niche while switching to and diversifying on conifers". Cladistics. 32 (5): 555–572. doi: 10.1111/cla.12141 . PMID   34740301. S2CID   86517289.
  14. Gullan, Penny J.; Martin, Jon H. (2009). "Sternorrhyncha". Encyclopedia of Insects (2nd ed.).
  15. Rohdendorf, B. B., ed. (1991). Fundamentals of Paleontology. Volume 9. Arthropoda, Tracheata, Chelicerata. Smithsonian Institution and National Science Foundation. pp. 267–274.
  16. Sorensen, J. T. (2009). "Aphids". In Resh, Vincent H.; Cardé, R. T. (eds.). Encyclopedia of Insects (2 ed.). Academic Press. pp. 27–31.
  17. "Superfamily Aphidoidea, Latreille, 1802". Aphid: species file. Archived from the original on 6 November 2017. Retrieved 3 February 2018.
  18. Blackman, R. L.; Eastrop, V. F. (1994). Aphids on the World's Trees. An Identification and Information Guide. Wallingford: CAB International. ISBN   978-0-85198-877-1.
  19. Granett, Jeffrey; Walker, M. Andrew; Kocsis, Laszlo; Omer, Amir D. (2001). "Biology and management of grape phylloxera". Annual Review of Entomology. 46: 387–412. doi:10.1146/annurev.ento.46.1.387. PMID   11112174.
  20. 1 2 3 4 5 6 McGavin, George C. (1993). Bugs of the World. Infobase Publishing. ISBN   978-0-8160-2737-8.
  21. Favret, C.; Eades, D.C. (2020). Miller, G.L.; Qiao, G.; Sano, Masakazu; Stekolshchikov, A.V. (eds.). "Aphid Species File - Aphidomorpha". Université de Montréal. aphid.speciesfile.org. Montreal, Canada: Aphid Species File. Archived from the original on February 24, 2021. Retrieved December 13, 2020.
  22. Papasotiropoulos, Vassilis; Tsiamis, Georgios; Papaioannou, Charikleia; Kilias, George (2013). "A molecular phylogenetic study of aphids (Hemiptera: Aphididae) based on mitochondrial DNA sequence analysis". Journal of Biological Research-Thessaloniki. 20: 1–13. Archived from the original on 2018-01-16. Retrieved 2018-01-15.
  23. Kim, Hyojoong; Lee, Seunghwan; Jang, Yikweon (September 2011). Moreau, Corrie S. (ed.). "Macroevolutionary Patterns in the Aphidini Aphids (Hemiptera: Aphididae): Diversification, Host Association, and Biogeographic Origins". PLOS ONE. 6 (9): e24749. Bibcode:2011PLoSO...624749K. doi: 10.1371/journal.pone.0024749 . PMC   3174202 . PMID   21935453.
  24. Ortiz-Rivas, Benjamín; Martínez-Torres, David (2010). "Combination of molecular data support the existence of three main lineages in the phylogeny of aphids (Hemiptera: Aphididae) and the basal position of the subfamily Lachninae". Molecular Phylogenetics and Evolution. 55 (1): 305–317. doi:10.1016/j.ympev.2009.12.005. PMID   20004730.
  25. Clark, Marta A; Moran, Nancy A.; Baumann, Paul; Wernegreen, Jennifer J. (2000). "Cospeciation Between Bacterial Endosymbionts (Buchnera) and a Recent Radiation of Aphids (Uroleucon) and Pitfalls of Testing for Phylogenetic Congruence". Evolution. 54 (2): 517–25. doi:10.1554/0014-3820(2000)054[0517:CBBEBA]2.0.CO;2. PMID   10937228.
  26. Nováková, Eva; Hypša, Václav; Klein, Joanne; Foottit, Robert G; von Dohlen, Carol D.; Moran, Nancy A. (2013). "Reconstructing the phylogeny of aphids (Hemiptera: Aphididae) using DNA of the obligate symbiont Buchnera aphidicola" (PDF). Molecular Phylogenetics and Evolution. 68 (1): 42–54. doi:10.1016/j.ympev.2013.03.016. PMID   23542003. Archived (PDF) from the original on 2017-08-10. Retrieved 2018-04-27.
  27. Chen, Rui; Wang, Zhe; Chen, Jing; Jiang, Li-Yun; Qiao, Ge-Xia (2017). "Insect-bacteria parallel evolution in multiple-co-obligate-aphid association: A case in Lachninae (Hemiptera: Aphididae)". Scientific Reports. 7 (1): 10204. Bibcode:2017NatSR...710204C. doi:10.1038/s41598-017-10761-9. PMC   5579299 . PMID   28860659.
  28. 1 2 3 4 5 6 7 8 Stroyan, Henry G. (1997). "Aphid". McGraw-Hill Encyclopedia of Science and Technology (8th ed.). ISBN   978-0-07-911504-1.
  29. Mutti, Navdeep S. (2006). Molecular Studies of the Salivary Glands of the Pea Aphid, Acyrthosiphon pisum (Harris) (PDF) (Ph.D. thesis). Kansas State University. Archived (PDF) from the original on 2012-02-27. Retrieved 2008-07-12.
  30. Jing, X.; White, T. A.; Yang, X.; Douglas, A. E. (2015). "The molecular correlates of organ loss: the case of insect Malpighian tubules - PMC". Biology Letters. 11 (5). doi:10.1098/rsbl.2015.0154. PMC   4455741 . PMID   25972400.
  31. 1 2 3 van Emden, Helmut F.; Harrington, Richard (2017). Aphids as Crop Pests. CABI. pp. 189–190. ISBN   978-1-78064-709-8.
  32. Powell, Glen; Tosh, Colin R.; Hardie, Jim (2005). "Host plant selection by aphids: Behavioral, Evolutionary, and Applied Perspectives". Annual Review of Entomology. 51 (1): 309–330. doi:10.1146/annurev.ento.51.110104.151107. PMID   16332214.
  33. Spiller, N. J.; Koenders, L.; Tjallingii, W. F. (1990). "Xylem ingestion by aphids – a strategy for maintaining water balance". Entomologia Experimentalis et Applicata. 55 (2): 101–104. doi:10.1007/BF00352570.
  34. 1 2 Fisher, D. B. (2000). "Long distance transport". In Buchanan, Bob B.; Gruissem, Wilhelm; Jones, Russell L. (eds.). Biochemistry and Molecular Biology of Plants (4th ed.). Rockville, Maryland: American Society of Plant Physiologists. pp.  730–784. ISBN   978-0-943088-39-6.
  35. Malone, M.; Watson, R.; Pritchard, J. (1999). "The spittlebug Philaenus spumarius feeds from mature xylem at the full hydraulic tension of the transpiration stream". New Phytologist. 143 (2): 261–271. doi: 10.1046/j.1469-8137.1999.00448.x . JSTOR   2588576.
  36. Powell, Glen; Hardie, Jim (2002). "Xylem ingestion by winged aphids". Entomologia Experimentalis et Applicata . 104 (1): 103–108. doi:10.1023/A:1021234412475.
  37. 1 2 3 4 Pompon, Julien; Quiring, Dan; Giordanengo, Philippe; Pelletier, Yvan (2010). "Role of xylem consumption on osmoregulation in Macrosiphum euphorbiae (Thomas)" (PDF). Journal of Insect Physiology . 56 (6): 610–615. doi:10.1016/j.jinsphys.2009.12.009. PMID   20036244. Archived from the original (PDF) on 16 July 2011.
  38. Kingsolver, J. G.; Daniel, T. L. (1995). "Mechanics of Food Handling by Fluid-Feeding Insects". In Chapman, R. F.; de Boer, Gerrit (eds.). Regulatory mechanisms in insect feeding. Springer. pp. 60–65.
  39. Ashford, D. A.; Smith, W. A.; Douglas, A. E. (2000). "Living on a high sugar diet: the fate of sucrose ingested by a phloem-feeding insect, the pea aphid Acyrthosiphon pisum". Journal of Insect Physiology . 46 (3): 335–341. doi:10.1016/S0022-1910(99)00186-9. PMID   12770238.
  40. Wilkinson, T. L.; Ashfors, D. A.; Pritchard, J.; Douglas, A. E. (1997). "Honeydew sugars and osmoregulation in the pea aphid Acyrthosiphon pisum". Journal of Experimental Biology. 200 (11): 2137–2143. doi:10.1242/jeb.200.15.2137. PMID   9320049. Archived from the original on 2008-08-21. Retrieved 2010-10-18.
  41. Shakesby, A. J.; Wallace, I. S.; Isaacs, H. V.; Pritchard, J.; Roberts, D. M.; Douglas, A. E. (2009). "A water-specific aquaporin involved in aphid osmoregulation". Insect Biochemistry and Molecular Biology. 39 (1): 1–10. doi:10.1016/j.ibmb.2008.08.008. PMID   18983920.
  42. Dadd, R. H.; Mittler, T. E. (1965). "Studies on the artificial feeding of the aphid Myzus persicae (Sulzer) – III. Some major nutritional requirements". Journal of Insect Physiology. 11 (6): 717–743. doi:10.1016/0022-1910(65)90154-X. PMID   5827534.
  43. Buchner, Paul (1965). Endosymbiosis of animals with plant microorganisms. Interscience. ISBN   978-0-470-11517-6.
  44. Whitehead, L. F.; Douglas, A. E. (1993). "A metabolic study of Buchnera, the intracellular bacterial symbionts of the pea aphid Acyrthosiphon pisum" (PDF). Journal of General Microbiology. 139 (4): 821–826. doi:10.1099/00221287-139-4-821.
  45. Febvay, Gérard; Liadouze, Isabelle; Guillaud, Josette; Bonnot, Guy (1995). "Analysis of energetic amino acid metabolism in Acyrthosiphon pisum: a multidimensional approach to amino acid metabolism in aphids". Archives of Insect Biochemistry and Physiology. 29 (1): 45–69. doi:10.1002/arch.940290106.
  46. Moran, Nancy A.; Jarvik, Tyler (2010). "Lateral transfer of genes from fungi underlies carotenoid production in aphids". Science . 328 (5978): 624–627. Bibcode:2010Sci...328..624M. doi:10.1126/science.1187113. PMID   20431015. S2CID   14785276.
  47. Altincicek, B.; Kovacs, J.L.; Gerardo, N.M. (2012). "Horizontally transferred fungal carotenoid genes in the two-spotted spider mite Tetranychus urticae". Biology Letters. 8 (2): 253–257. doi:10.1098/rsbl.2011.0704. PMC   3297373 . PMID   21920958.
  48. Valmalette, Jean Christophe; Dombrovsky, Aviv; Brat, Pierre; Mertz, Christian; Capovilla, Maria; Robichon, Alain (2012). "Light-induced electron transfer and ATP synthesis in a carotene synthesizing insect". Scientific Reports. 2: 579. Bibcode:2012NatSR...2E.579V. doi:10.1038/srep00579. PMC   3420219 . PMID   22900140.
  49. 1 2 Wang, C. L.; Siang, L. Y.; Chang, G. S.; Chu, H. F. (1962). "Studies on the soybean aphid, Aphis glycines Matsumura". Acta Entomologica Sinica. 11: 31–44.
  50. 1 2 van Emden, Helmut F.; Harrington, Richard (2017). Aphids as Crop Pests, 2nd Edition. CABI. pp. 81–82. ISBN   978-1-78064-709-8. Archived from the original on 2021-06-24. Retrieved 2018-04-29.
  51. Von Dohlen, Carol; Moran, Nancy A. (2000). "Molecular data support a rapid radiation of aphids in the Cretaceous and multiple origins of host alternation". Biological Journal of the Linnean Society. 71 (4): 689–717. doi:10.1006/bijl.2000.0470.
  52. Moran, Nancy A. (1992). "The Evolution of Aphid Life Cycles". Annual Review of Entomology. 37: 321–348. doi:10.1146/annurev.en.37.010192.001541.
  53. 1 2 Blackman, Roger L. (2008). "Stability and variation in aphid clonal lineages". Biological Journal of the Linnean Society. 11 (3): 259–277. doi:10.1111/j.1095-8312.1979.tb00038.x.
  54. 1 2 Gullan, P. J.; Cranston, P. S. (2010). The Insects: An Outline of Entomology (4th ed.). Wiley. pp. 150–151. ISBN   978-1-118-84615-5.
  55. 1 2 3 Hales, Dinah F.; Wilson, Alex C. C.; Sloane, Mathew A.; Simon, Jean-Christophe; Legallic, Jean-François; Sunnucks, Paul (2002). "Lack of Detectable Genetic Recombination on the X Chromosome During the Parthenogenetic Production of Female and Male Aphids". Genetics Research. 79 (3): 203–209. doi: 10.1017/S0016672302005657 . PMID   12220127. S2CID   25605477.
  56. Nevo, Ettay; Coll, Moshe (2001). "Effect of nitrogen fertilization on Aphis gossypii (Homoptera: Aphididae): variation in size, color, and reproduction". Journal of Economic Entomology . 94 (1): 27–32. doi: 10.1603/0022-0493-94.1.27 . PMID   11233124. S2CID   25758038.
  57. Jahn, Gary C.; Almazan, Liberty P.; Pacia, Jocelyn B. (2005). "Effect of nitrogen fertilizer on the intrinsic rate of increase of the rusty plum aphid, Hysteroneura setariae (Thomas) (Homoptera: Aphididae) on rice (Oryza sativa L.)" (PDF). Environmental Entomology . 34 (4): 938–943. doi:10.1603/0046-225X-34.4.938. S2CID   1941852. Archived from the original (PDF) on 2010-09-09.
  58. Hughes, R. D. (1963). "Population Dynamics of the Cabbage Aphid, Brevicoryne brassicae (L.)". Journal of Animal Ecology. 32 (3): 393–424. doi:10.2307/2600. JSTOR   2600.
  59. Brisson, Jennifer A. (2010). "Aphid wing dimorphisms: linking environmental and genetic control of trait variation". Philosophical Transactions of the Royal Society B: Biological Sciences. 365 (1540): 605–616. doi:10.1098/rstb.2009.0255. PMC   2817143 . PMID   20083636.
  60. Weisser, Wolfgang W.; Zytynska, Sharon E.; Mehrparvar, Mohsen (2013-03-05). "Multiple Cues for Winged Morph Production in an Aphid Metacommunity". PLOS ONE. 8 (3): e58323. Bibcode:2013PLoSO...858323M. doi: 10.1371/journal.pone.0058323 . ISSN   1932-6203. PMC   3589340 . PMID   23472179.
  61. Lees, A. D. (1967-02-01). "The production of the apterous and alate forms in the aphid Megoura viciae Buckton, with special reference to the rôle of crowding". Journal of Insect Physiology. 13 (2): 289–318. doi:10.1016/0022-1910(67)90155-2. ISSN   0022-1910.
  62. Kunert, Grit; Otto, Susanne; Röse, Ursula S. R.; Gershenzon, Jonathan; Weisser, Wolfgang W. (2005-04-28). "Alarm pheromone mediates production of winged dispersal morphs in aphids". Ecology Letters. 8 (6): 596–603. doi:10.1111/j.1461-0248.2005.00754.x. ISSN   1461-023X.
  63. Ryabov, E. V.; Keane, G.; Naish, N.; Evered, C.; Winstanley, D. (2009-05-13). "Densovirus induces winged morphs in asexual clones of the rosy apple aphid, Dysaphis plantaginea". Proceedings of the National Academy of Sciences. 106 (21): 8465–8470. Bibcode:2009PNAS..106.8465R. doi: 10.1073/pnas.0901389106 . ISSN   0027-8424. PMC   2688996 . PMID   19439653.
  64. Chan, C. K. (1991). Aphid-transmitted viruses and their vectors of the world . Research Branch, Agriculture Canada. ISBN   0662183347. OCLC   872604083.
  65. Reyes, Miguel L.; Laughton, Alice M.; Parker, Benjamin J.; Wichmann, Hannah; Fan, Maretta; Sok, Daniel; Hrček, Jan; Acevedo, Tarik; Gerardo, Nicole M. (2019-01-31). "The influence of symbiotic bacteria on reproductive strategies and wing polyphenism in pea aphids responding to stress". Journal of Animal Ecology. 88 (4): 601–611. doi:10.1111/1365-2656.12942. ISSN   0021-8790. PMC   6453707 . PMID   30629747.
  66. Alford, David V. (2014). Pests of Fruit Crops: A Colour Handbook, Second Edition. CRC Press. pp. 71–72. ISBN   978-1-4822-5421-1. Archived from the original on 2021-06-24. Retrieved 2018-02-03.
  67. 1 2 Hooper-Bui, Linda M. (2008). "Ant". World Book Encyclopedia . ISBN   978-0-7166-0108-1.
  68. Stadler, Bernhard; Dixon, Anthony F. G. (2005). "Ecology and Evolution of Aphid-Ant Interactions". Annual Review of Ecology, Evolution, and Systematics. 36 (1): 345–372. doi:10.1146/annurev.ecolsys.36.091704.175531.
  69. Wootton, Anthony (1998). Insects of the World. Blandford. ISBN   978-0-7137-2366-3.
  70. 1 2 Neary, John (1977). Insects and Spiders. Time-Life Books. pp.  78–79. ISBN   978-0-8094-9687-7.
  71. Salazar, Adrián; Fürstenau, Benjamin; Quero, Carmen; Pérez-Hidalgo, Nicolás; Carazo, Pau; Font, Enrique; Martínez-Torres, David (2015). "Aggressive mimicry coexists with mutualism in an aphid". Proceedings of the National Academy of Sciences. 112 (4): 1101–1106. Bibcode:2015PNAS..112.1101S. doi: 10.1073/pnas.1414061112 . PMC   4313836 . PMID   25583474.
  72. Baumann, Paul; Moran, Nancy A.; Baumann, Linda (2006). "Bacteriocyte-Associated Endosymbionts of Insects". In Dworkin, Martin; Falkow, Stanley; Rosenberg, Eugene; Schleifer, Karl-Heinz; Stackebrandt, Erko (eds.). The Prokaryotes. pp. 403–438. doi:10.1007/0-387-30741-9_16. ISBN   978-0-387-25476-0.
  73. Douglas, A. E. (1998). "Nutritional Interactions in Insect-Microbial Symbioses: Aphids and Their Symbiotic BacteriaBuchnera". Annual Review of Entomology. 43 (1): 17–37. doi:10.1146/annurev.ento.43.1.17. PMID   15012383.
  74. Bodył, Andrzej; Mackiewicz, Paweł; Gagat, Przemysław (2012). "Organelle Evolution: Paulinella Breaks a Paradigm". Current Biology. 22 (9): R304–R306. doi: 10.1016/j.cub.2012.03.020 . PMID   22575468.
  75. Baumann, Linda; Baumann, Paul; Moran, Nancy A.; Sandström, Jonas; Thao, Mylo Ly (January 1999). "Genetic Characterization of Plasmids Containing Genes Encoding Enzymes of Leucine Biosynthesis in Endosymbionts (Buchnera) of Aphids". Journal of Molecular Evolution. 48 (1): 77–85. Bibcode:1999JMolE..48...77B. doi:10.1007/PL00006447. ISSN   0022-2844. PMID   9873079. S2CID   24062989. Archived from the original on 2021-11-13. Retrieved 2020-11-16.
  76. Sabri, Ahmed; Vandermoten, Sophie; Leroy, Pascal D.; Haubruge, Eric; Hance, Thierry; Thonart, Philippe; De Pauw, Edwin; Francis, Frédéric (2013-09-25). "Proteomic Investigation of Aphid Honeydew Reveals an Unexpected Diversity of Proteins". PLOS ONE. 8 (9): e74656. Bibcode:2013PLoSO...874656S. doi: 10.1371/journal.pone.0074656 . ISSN   1932-6203. PMC   3783439 . PMID   24086359.
  77. Perez-Brocal, V.; Gil, R.; Ramos, S.; Lamelas, A.; Postigo, M.; Michelena, J.M.; Silva, F. J.; Moya, A.; Latorre, A. (2006). "A Small Microbial Genome: The End of a Long Symbiotic Relationship?". Science. 314 (5797): 312–313. Bibcode:2006Sci...314..312P. doi:10.1126/science.1130441. PMID   17038625. S2CID   40081627.
  78. Mira, A.; Moran, Nancy A. (2002). "Estimating Population Size and Transmission Bottlenecks in Maternally Transmitted Endosymbiotic Bacteria". Microbial Ecology. 44 (2): 137–143. doi:10.1007/s00248-002-0012-9. PMID   12087426. S2CID   33681686.
  79. Sakaki, Yoshiyuki; Shigenobu, Shuji; Watanabe, Hidemi; Hattori, Masahira; Ishikawa, Hajime (2000). "Genome sequence of the endocellular bacterial symbiont of aphids Buchnera sp. APS". Nature. 407 (6800): 81–86. Bibcode:2000Natur.407...81S. doi: 10.1038/35024074 . PMID   10993077.
  80. Viñuelas, José; Calevro, Federica; Remond, Didier; Bernillon, Jacques; Rahbé, Yvan; Febvay, Gérard; Fayard, Jean-Michel; Charles, Hubert (2007). "Conservation of the links between gene transcription and chromosomal organization in the highly reduced genome of Buchnera aphidicola". BMC Genomics. 8 (1): 143. doi: 10.1186/1471-2164-8-143 . PMC   1899503 . PMID   17547756.
  81. Moran, Nancy A.; Dunbar, Helen E.; Wilcox, Jennifer L. (2005). "Regulation of Transcription in a Reduced Bacterial Genome: Nutrient-Provisioning Genes of the Obligate Symbiont Buchnera aphidicola". Journal of Bacteriology. 187 (12): 4229–4237. doi:10.1128/JB.187.12.4229-4237.2005. PMC   1151715 . PMID   15937185.
  82. Tsuchida, T.; Koga, R.; Meng, X. Y.; T. Matsumoto; T. Fukatsu (2005). "Characterization of a facultative endosymbiotic bacterium of the pea aphid Acyrthosiphon pisum". Microbial Ecology . 49 (1): 126–133. doi:10.1007/s00248-004-0216-2. PMID   15690225. S2CID   24144752.
  83. Sakurai, M.; Koga, R.; Tsuchida, T.; Meng, X.-Y.; Fukatsu, T. (2005). "Rickettsia Symbiont in the Pea Aphid Acyrthosiphon pisum: Novel Cellular Tropism, Effect on Host Fitness, and Interaction with the Essential Symbiont Buchnera". Applied and Environmental Microbiology. 71 (7): 4069–4075. Bibcode:2005ApEnM..71.4069S. doi:10.1128/AEM.71.7.4069-4075.2005. PMC   1168972 . PMID   16000822.
  84. Ferrari, Julia; Scarborough, Claire L.; Godfray, H. Charles J. (2007). "Genetic variation in the effect of a facultative symbiont on host-plant use by pea aphids". Oecologia. 153 (2): 323–329. Bibcode:2007Oecol.153..323F. doi:10.1007/s00442-007-0730-2. PMID   17415589. S2CID   37052892.
  85. Simon, J.-C.; Carre, S.; Boutin, M.; Prunier-Leterme, N.; Sabater-Munoz, B.; Latorre, A.; Bournoville, R. (2003). "Host-based divergence in populations of the pea aphid: insights from nuclear markers and the prevalence of facultative symbionts". Proceedings of the Royal Society B: Biological Sciences. 270 (1525): 1703–1712. doi:10.1098/rspb.2003.2430. PMC   1691435 . PMID   12964998.
  86. Weldon, Stephanie R.; Oliver, Kerry M. (2016). "Diverse Bacteriophage Roles in an Aphid-Bacterial Defensive Mutualism". The Mechanistic Benefits of Microbial Symbionts. Advances in Environmental Microbiology. Vol. 2. Springer, Cham. pp. 173–206. doi:10.1007/978-3-319-28068-4_7. ISBN   9783319280660.
  87. Weldon, S. R.; Strand, M. R.; Oliver, K. M. (2013-01-22). "Phage loss and the breakdown of a defensive symbiosis in aphids". Proceedings of the Royal Society of London B: Biological Sciences. 280 (1751): 20122103. doi:10.1098/rspb.2012.2103. PMC   3574403 . PMID   23193123.
  88. Oliver, K. M.; Moran, Nancy A.; Hunter, M. S. (2006). "Costs and benefits of a superinfection of facultative symbionts in aphids". Proceedings of the Royal Society B: Biological Sciences. 273 (1591): 1273–1280. doi:10.1098/rspb.2005.3436. PMC   1560284 . PMID   16720402.
  89. 1 2 Capinera, John (2011). Insects and Wildlife: Arthropods and their Relationships with Wild Vertebrate Animals. John Wiley & Sons. p. 536. ISBN   978-1-4443-5784-4.
  90. van Emden, Helmut F.; Harrington, Richard (2017). Aphids as Crop Pests. CABI. pp. 229–230. ISBN   978-1-78064-709-8.
  91. 1 2 3 Brust, Gerald E. (22 June 2006). "Early season aphid and thrips populations". University of Maryland, College Park. Archived from the original on 19 July 2011. Retrieved 18 October 2010.
  92. Lamb, K. P. (1961). "Some effects of fluctuating temperatures on metabolism, development, and rate of population growth in the cabbage aphid, Brevicoryne brassicae". Ecology . 42 (4): 740–745. doi:10.2307/1933502. JSTOR   1933502.
  93. Jones, Margaret G. (1979). "Abundance of aphids on cereals from before 1973 to 1977". Journal of Applied Ecology . 16 (1): 1–22. doi:10.2307/2402724. JSTOR   2402724.
  94. Krupke, Christian; Obermeyer, John; O'Neil, Robert (2007). "Soybean aphid, a new beginning for 2007". Pest and Crop. Purdue University. 7. Archived from the original on 2021-02-04. Retrieved 2008-07-17.
  95. "Why some aphids can't stand the heat". Science Daily . 23 April 2007. Archived from the original on 15 January 2018. Retrieved 28 February 2018.
  96. Hughes, R.D. (1963). "Population dynamics of the cabbage aphid, Brevicoryne brassicae (L.)". Journal of Animal Ecology . 32 (3): 393–424. doi:10.2307/2600. JSTOR   2600.
  97. Suwanbutr, S. (1996). "Stable age distributions of lucerne aphid populations in SE-Tasmania" (PDF). Thammasat International Journal of Science and Technology. 1 (5): 38–43. Archived from the original (PDF) on 10 September 2008.
  98. Ostlie, Ken (3 August 2006). "Spider Mites, Aphids and Rain Complicating Spray Decisions in Soybean" (PDF). Minnesota Crop eNews. University of Minnesota. Archived from the original (PDF) on 10 September 2008.
  99. Aoki, S. (1977). "Colophina clematis (Homoptera, Pemphigidae), an aphid species with soldiers" (PDF). Japanese Journal of Entomology . 45 (2): 276–282.[ permanent dead link ]
  100. 1 2 3 Kutsukake, M.; Meng, X.Y.; Katayama, N.; Nikoh, N.; Shibao, H.; Fukatsu, T. (2012). "An insect-induced novel plant phenotype for sustaining social life in a closed system". Nature Communications. 3: 1187–1193. Bibcode:2012NatCo...3.1187K. doi:10.1038/ncomms2187. PMC   3514493 . PMID   23149732.
  101. Preston-Mafham, Rod; Preston-Mafham, Ken (1993). The Encyclopedia of Land Invertebrate Behaviour . MIT Press. p.  281. ISBN   978-0-262-16137-4.
  102. Gish, M.; Dafni, A.; Inbar, M. (2012). Heil, Martin (ed.). "Young Aphids Avoid Erroneous Dropping when Evading Mammalian Herbivores by Combining Input from Two Sensory Modalities". PLOS ONE. 7 (4): e32706. Bibcode:2012PLoSO...732706G. doi: 10.1371/journal.pone.0032706 . PMC   3322135 . PMID   22496734.
  103. Kazana, Eleanna; Pope, Tom W.; Tibbles, Laurienne; Bridges, Matthew; Pickett, John A.; Bones, Atle M.; Powell, Glen; Rossiter, John T. (2007). "The cabbage aphid: a walking mustard oil bomb". Proceedings of the Royal Society B. 274 (1623): 2271–2277. doi:10.1098/rspb.2007.0237. PMC   2288485 . PMID   17623639.
  104. Vilcinskas, Andreas (2016). "Aphid Immunity.". Biology and Ecology of Aphids. CRC Press. p. 131.
  105. Edwards, John S. (1966). "Defence by smear: supercooling in the cornicle wax of aphids". Nature . 211 (5044): 73–74. Bibcode:1966Natur.211...73E. doi:10.1038/211073a0. S2CID   4295930.
  106. Martinson, Candace (2008). "Aphid". World Book Encyclopedia . ISBN   978-0-7166-0108-1.
  107. Way, M.J. (1963). "Mutualism between ants and honeydew-producing Homoptera". Annual Review of Entomology . 8: 307–344. doi:10.1146/annurev.en.08.010163.001515.
  108. Grasswitz, Tessa R.; Paine, Timothy D. (1992). "Kairomonal effect of an aphid cornicle secretion on Lysiphlebus testaceipes (Cresson) (Hymenoptera: Aphidiidae)". Journal of Insect Behavior . 5 (4): 447–457. doi:10.1007/BF01058190. S2CID   25298742.
  109. Lenhart, Paul A.; Jackson, Kelly A.; White, Jennifer A. (2018). "Heritable variation in prey defence provides refuge for subdominant predators". Proceedings of the Royal Society B: Biological Sciences. 285 (1879): 20180523. doi:10.1098/rspb.2018.0523. PMC   5998095 . PMID   29848647.
  110. "Aphid Parasitoids". IET Department of the College of Agriculture and Natural Resources. extension.umd.edu. College Park, Maryland: University of Maryland. Archived from the original on 2018-06-26. Retrieved 2018-04-29 via College of Agriculture & Natural Sciences. Aphid parasitoids are very tiny wasps, about 1/10 inch long. They are slender, black or brown, and have a pinched or "wasp waist"
  111. Havelka, Jan; Tomanović, Željko; Kavallieratos, Nickolas G.; Rakhshani, Ehsan; Pons, Xavier; Petrović, Andjeljko; Pike, Keith S.; Starý, Petr (2012-05-01). "Review and Key to the World Parasitoids (Hymenoptera: Braconidae: Aphidiinae) of Aphis ruborum (Hemiptera: Aphididae) and Its Role as a Host Reservoir". Annals of the Entomological Society of America. 105 (3): 386–394. doi: 10.1603/an11108 . S2CID   84348019.
  112. 1 2 3 Flint, M.L. (July 2013). "Aphids". UC IPM. Archived from the original on 2018-04-09. Retrieved 6 February 2018.
  113. Gibson, R. W.; Pickett, J. A. (1983). "Wild potato repels aphids by release of aphid alarm pheromone". Nature. 302 (5909): 608–609. Bibcode:1983Natur.302..608G. doi:10.1038/302608a0. S2CID   4345998.
  114. Gibson, R. W. (1971). "Glandular hairs providing resistance to aphids in certain wild potato species". Annals of Applied Biology. 68 (2): 113–119. doi:10.1111/j.1744-7348.1971.tb06448.x.
  115. United States Department of Agriculture (2003). Studies on the Chemical and Biological Properties of Coumestrol and Related Compounds. US Government Printing Office. pp. 47–67. Archived from the original on 2015-04-02. Retrieved 2015-04-26.
  116. Gillman, Daniel H. (2005). "Sooty mold" (PDF). University of Massachusetts Amherst. Archived (PDF) from the original on September 28, 2011. Retrieved October 18, 2010.
  117. Reynolds, Hannah T.; Volk, Tom (September 2007). "Scorias spongiosa, the beech aphid poop-eater". Tom Volk's Fungus of the Month. University of Wisconsin–La Crosse. Archived from the original on July 30, 2010. Retrieved October 18, 2010.
  118. Dik, A. J.; van Pelt, J. A. (1992). "Interaction between phyllosphere yeasts, aphid honeydew and fungicide effectiveness in wheat under field conditions". Plant Pathology . 41 (6): 661–675. doi:10.1111/j.1365-3059.1992.tb02550.x.
  119. Choudhury, Dhrupad (1985). "Aphid honeydew: a re-appraisal of the hypothesis of Owen and Wiegert". Oikos . 45 (2): 287–290. doi:10.2307/3565718. JSTOR   3565718.
  120. Stern, D. L.; Foster, W. A. (1996). "The evolution of soldiers in aphids". Biological Reviews of the Cambridge Philosophical Society. 71 (1): 27–79. doi:10.1111/j.1469-185X.1996.tb00741.x. PMID   8603120. S2CID   8991755.
  121. Choe; Jae C.; Crespi, Bernard J. (1997). The Evolution of Social Behaviour in Insects and Arachnids. Cambridge University Press. pp. 150–152. ISBN   978-0-521-58977-2.{{cite book}}: CS1 maint: multiple names: authors list (link)
  122. Pike, N.; Richard, D.; Foster, W.; Mahadevan, L. (2002). "How aphids lose their marbles". Proceedings of the Royal Society B: Biological Sciences. 269 (1497): 1211–5. doi:10.1098/rspb.2002.1999. PMC   1691028 . PMID   12065036.
  123. 1 2 van Emden, Helmut F.; Harrington, Richard (2017). Aphids as Crop Pests. CABI. pp. 196–198. ISBN   978-1-78064-709-8.
  124. Cloyd, Raymond (14 February 2022). "Aphids". Greenhouse Product News. 32: 14.
  125. Shulman, Sidney (1967-01-01). "Allergic Responses to Insects". Annual Review of Entomology. 12 (1): 323–346. doi:10.1146/annurev.en.12.010167.001543.
  126. Pundt, Leanne (2011). "Managing Aphids in the Greenhouse". University of Connecticut College of Agriculture, Health, and Natural Resources. Archived from the original on 2018-02-18. Retrieved 20 February 2018.
  127. Cranshaw, W.S. (March 2008). "Insect Control: Soaps and Detergents". University of Colorado. Archived from the original on 7 February 2018. Retrieved 6 February 2018.
  128. Ubl, J.B. (July 2009). "Insecticidal Soaps for Garden Pest Control". Clemson University. Archived from the original on 22 January 2018. Retrieved 6 February 2018.
  129. Evans, D. A.; Medler, J. T. (1966-12-01). "Improved Method of Using Yellow-Pan Aphid Traps1". Journal of Economic Entomology. 59 (6): 1526–1527. doi:10.1093/jee/59.6.1526. ISSN   1938-291X. Archived from the original on 2021-11-13. Retrieved 2020-03-22.
  130. Döring, Thomas Felix; Chittka, Lars (2007). "Visual ecology of aphids—a critical review on the role of colours in host finding" (PDF). Arthropod-Plant Interactions. 1 (1): 3–16. doi:10.1007/s11829-006-9000-1. S2CID   20066025. Archived (PDF) from the original on 2020-09-19. Retrieved 2020-08-27.
  131. Jaronski, S. T.; Mascarin, G. M. (2017). Lacey, Lawrence (ed.). Mass Production of Fungal Entomopathogens. Academic Press. pp. 141–155. ISBN   978-0-12-803527-6.{{cite book}}: |work= ignored (help)
  132. Steinkraus, Donald C. (2006). "Factors affecting transmission of fungal pathogens of aphids". Journal of Invertebrate Pathology. 92 (3): 125–131. doi:10.1016/j.jip.2006.03.009. PMID   16780867.
  133. Lady Beetle Releases for Aphid Control: How to Help Them Work Archived 2015-04-02 at the Wayback Machine . Clark, J.K., University of California Davis, June 2011.
  134. Beale, M. H.; Birkett, M. A.; Bruce, T. J. A.; Chamberlain, K.; Field, L. M.; Huttly, A. K.; Martin, J. L.; Parker, R.; Phillips, A. L.; Pickett, J. A.; Prosser, I. M.; Shewry, P. R.; Smart, L. E.; Wadhams, L. J.; Woodcock, C. M.; Zhang, Y. (2006). "Aphid alarm pheromone produced by transgenic plants affects aphid and parasitoid behavior". Proceedings of the National Academy of Sciences. 103 (27): 10509–10513. Bibcode:2006PNAS..10310509B. doi: 10.1073/pnas.0603998103 . PMC   1502488 . PMID   16798877.
  135. Pickett, J. A.; Wadhams, L. J.; Woodcock, C. M.; Hardie, J. (1992). "The Chemical Ecology of Aphids". Annual Review of Entomology. 37 (1): 67–90. doi:10.1146/annurev.en.37.010192.000435.
  136. Marren, Peter; Mabey, Richard (2010). Bugs Britannica. Chatto & Windus. pp. 191–194. ISBN   978-0-7011-8180-2. Archived from the original on 2017-03-14. Retrieved 2018-02-09.

On the University of Florida / Institute of Food and Agricultural Sciences Featured Creatures website: