ChapterPDF Available

Biodiversity of Diptera

Authors:

Abstract and Figures

Overview of TaxaSocietal ImportanceReferences
Content may be subject to copyright.
Chapter 9
Biodiversity
of Diptera
Gregory W. Courtney1, Thomas Pape2,
Jeffrey H. Skevington3, and Bradley J. Sinclair4
1Department of Entomology, 432 Science II, Iowa State University, Ames, Iowa
50011 USA
2Natural History Museum of Denmark, Zoological Museum, Universitetsparken 15,
DK 2100 Copenhagen Denmark
3Agriculture and Agri-Food Canada, Canadian National Collection of Insects, Arachnids
and Nematodes, K.W. Neatby Building, 960 Carling Avenue, Ottawa, Ontario K1A 0C6
Canada
4Entomology Ontario Plant Laboratories, Canadian Food Inspection Agency, K.W.
Neatby Building, 960 Carling Avenue, Ottawa, Ontario K1A 0C6 Canada
Insect Biodiversity: Science and Society, 1st edition. Edited by R. Foottit and P. Adler
©2009 Blackwell Publishing, ISBN 978-1-4051-5142-9
185
The Diptera, commonly called true flies or
two-winged flies, are a familiar group of
insects that includes, among many others,
black flies, fruit flies, horse flies, house flies, midges,
and mosquitoes. The Diptera are among the most
diverse insect orders, with estimates of described
richness ranging from 120,000 to 150,000 species
(Colless and McAlpine 1991, Schumann 1992, Brown
2001, Merritt et al. 2003). Our world tally of more
than 152,000 described species (Table 9.1) is based
primarily on figures extracted from the ‘BioSystematic
Database of World Diptera’ (Evenhuis et al. 2007).
The Diptera are diverse not only in species richness,
but also in their structure (Fig. 9.1), habitat exploita-
tion, life habits, and interactions with humankind
(Hennig 1973, McAlpine et al. 1981, 1987, Papp
and Darvas 2000, Brown 2001, Skevington and Dang
2002, Pape 2009). The Diptera have successfully
colonized all continents, including Antarctica, and
practically every habitat except the open sea and
inside glaciers. Larval Diptera are legless (Figs. 9.2 and
9.3F-J) and found in a variety of terrestrial and aquatic
habitats (Teskey 1976, Ferrar 1987, H ¨ovemeyer
2000, Courtney and Merritt 2008). Larvae of most
species can be considered aquatic in the broadest
sense because, for survival, they require a moist to wet
environment within the tissues of living plants, amid
decaying organic materials, as parasites or parasitoids
of other animals, or in association with bodies of water.
Most larvae are free-living and swim, crawl, or tunnel
actively in water (e.g., Chaoboridae, Chironomidae,
Culicidae, and Simuliidae), sediments (e.g., Cerato-
pogonidae, Psychodidae, Tabanidae, and Tipulidae),
wood (e.g., Axymyiidae, some Syrphidae and
Tipulidae), fruit (e.g., Chloropidae and Tephritidae), or
decaying organic material (e.g., Ephydridae, Muscidae,
Sarcophagidae, and Sphaeroceridae). Other larvae
inhabit the tissues of living organisms (e.g., Acroceri-
dae, Oestridae, Pipunculidae, and Tachinidae). Still
others (e.g., larvae of the superfamily Hippoboscoidea)
are retained and nourished in the female abdomen
until deposited and ready to quickly pupariate. Most of
the feeding and accumulation of biomass occurs in the
larval stage, and adult Diptera mostly take only what
they need to supply their flight muscles with energy.
Among those flies that feed extensively, their diets con-
sist of nectar or honeydew (e.g., Blephariceridae and
Bombyliidae), pollen (e.g., Nemestrinidae and Syrphi-
dae), vertebrate blood (e.g., Culicidae and Glossinidae),
insect hemolymph (e.g., some Ceratopogonidae), and
other organic materials that are liquified or can be
dissolved or suspended in saliva or regurgitated fluid
(e.g., Calliphoridae, Micropezidae, and Muscidae). The
adults of some groups are predaceous (e.g., Asilidae,
Empididae, and some Scathophagidae), whereas those
of a few Diptera (e.g., Deuterophlebiidae and Oestridae)
lack mouthparts completely, do not feed, and live for
onlyabrieftime.
As holometabolous insects that undergo complete
metamorphosis, the Diptera have a life cycle that
includes a series of distinct stages or instars. A typi-
cal life cycle consists of a brief egg stage (usually a few
days or weeks, but sometimes much longer), three or
four larval instars (usually three in Brachycera, four in
lower Diptera, more in Simuliidae, Tabanidae, Thau-
maleidae, and a few others), a pupal stage of varying
length, and an adult stage lasting from less than 2 h
(Deuterophlebiidae) to several weeks or even months.
The eggs of Diptera are laid singly, in small clusters, or
in loose or compact masses, and they can be attached
to rocks, vegetation, or other substrata, or deposited
on or in the food source. Oviposition sites are usu-
ally in or near the larval habitat, which ensures that
eggs are placed in a location suitable for larval devel-
opment, with a notable exception being the human
bot fly, Dermatobia hominis, which glues its eggs to
zoophilous dipterans (e.g., calyptrate flies and Culi-
cidae), thereby ensuring a carrier-mediated infection
(Guimar˜aes and Papavero 1999). In some groups, eggs
are incubated and hatch during (e.g., Sarcophagidae) or
immediately after deposition (e.g., many Tachinidae),
or the female is truly viviparous when the larvae are
nourished and grow while still inside the female (e.g.,
Hippoboscoidea, and mesembrinelline Calliphoridae)
(Ferrar 1987, Meier et al. 1999). For a given species,
all larval instars usually occur in the same habitat.
In general, the duration of the first instar is shortest,
whereas that of the last instar is much longer, often
several weeks or even months. Although most Diptera
exhibit sexual reproduction, parthenogenesis occurs in
some groups, and reproduction by immature stages
(paedogenesis) has been recorded in some gall midges
(Cecidomyiidae).
Among the most unusual life histories is that
of the Nymphomyiidae. Adults have a larviform
appearance, lack mouthparts (Fig. 9.3B), and possess
wings that are deciduous, elongate, and fringed with
long microtrichia (Courtney 1994). Most species
are associated with small headwater streams where
larvae, pupae, and copulating adults occur on rocks
Biodiversity of Diptera 187
covered with aquatic mosses. Although few details
about mating behavior are available, observations of
Appalachian species suggest that adults locate a mate
soon after emergence, couple, descend into the water
in copula, shed their wings, and crawl to an oviposition
site. The female then lays a rosette of eggs around the
coupled adults, which die in copula (Courtney 1994).
Another remarkable life history is that of Fergusonina
turneri (Fergusoninidae), which in an obligate mutu-
alism with the nematode Fergusobia quinquenerviae is
gall building on the myrtacean plant Melaleuca quin-
quenervia (Taylor 2004). Galls are initiated in buds and
young leaves by juvenile nematodes, which are injected
by ovipositing female flies, along with their own eggs.
When the fly eggs hatch, the larvae form individual
cavities in the galls, and nematodes move into these
and coexist with the fly larva. The nematodes pass
through at least one parthenogenetic generation, and
fertilized female nematodes of a later sexual generation
invade the late third-instar fly larva. Nematode eggs are
deposited in the larval hemolymph and, after hatching,
the juvenile nematodes migrate to the fly ovaries. When
the adult female fly hatches, it will continue the cycle
by depositing new nematodes along with its own eggs.
In the Phoridae, the peculiarly swollen, physogastric
females of species in the subfamily Termitoxeniinae,
which are all associated with fungus-growing termites,
show a post-metamorphic growth in both head and
hind legs, which is unique for an adult, nonmolt-
ing insect (Disney and Kistner 1995). These termite
inquilines were described by Wasmann (1910: 38) as
‘a store-house of anomalies, whether we consider them
from the point of view of morphologists, anatomists,
evolutionists, or biologists. They are exceptions to the
laws of entomology’.
Some predaceous species of flies have evolved odd
larval lifestyles. Adults of Oedoparena glauca (Dryomyzi-
dae) oviposit on closed barnacles during low tide. The
eggs hatch during subsequent low-tide periods and lar-
vae enter the barnacles as they open, with the incoming
tide. During high tide, larvae feed inside the tissues of
the submerged barnacles, and in subsequent low-tide
periods they search for new prey (Burger et al. 1980).
The larvae of the Vermileonidae are commonly called
worm lions because they construct pitfall traps similar
to those of ant lions (family Myrmeleontidae) of the
order Neuroptera. The worm lion waits buried in the
bottom of the pit for an insect prey to tumble in, pounces,
sucks out its body juices, and then tosses the victim’s
corpse from the pit (Wheeler 1930, Teskey 1981). A
number of species in the Keroplatidae (Orfelia fultoni,
Arachnocampa spp., and Keroplatus spp.) are biolumi-
nescent and emit a blue-green light as larvae. These
glowworms construct mucous tubes from which they
hang snares with droplets of oxalic acid that capture
and kill prey attracted by their bioluminescence. The
larvae are voracious predators and feed on many types
of arthropods attracted to their glow (Baker 2002). In
some instances, these glowworms congregate in large
numbers and form impressive displays. For example, a
superb concentration of Arachnocampa luminosa in the
Waitoma Caves in New Zealand attracts more than
300,000 visitors per year (Baker 2002).
A common mating behavior among the lower
Diptera (Chaoboridae, Chironomidae, and others)
is the formation of dense and sometimes enormous
swarms (Vockeroth 2002). The swarms generally are
composed of males, and when females enter the swarm,
coupling quickly takes place. Males often exhibit
adaptations that enable mate detection. For example,
the eyes of males of most species engaged in swarming
are enlarged and contiguous above (presumably to
aid in spotting females from below) and the antennae
have numerous, long, hairlike setae that allow them
to detect a female’s wing beats. In the Brachycera,
premating behavior includes posturing and displays
in courtship that can become complex performances
with combinations of kneeling, jumping, and flapping
(e.g., Struwe 2005). In dance flies (Empididae), which
often have mating swarms, the male presents the
female with an edible lure or an inedible substitute to
initiate mating (Cumming 1994).
Many Diptera congregate at landmarks for the pur-
pose of mating. Landmarks can range from a rock to
a tuft of grass, a road, a stream course, a canyon, a
bog, an emergent tree (taller than the others), or a
hilltop. The difference between simple landmarks and
hilltops is that simple landmarks typically support only
a single species. However, emergent trees in rainforests
are likely immensely important for landmark mat-
ing species, although few data are available. Hilltops
are significant landmarks because they support many
species, often hundreds or, in rare cases, even thou-
sands (Skevington 2008). Hilltops range from massive,
rocky mountaintops more than 4000 m high to small
hummocks in flat country. The height above the sur-
rounding land must not be too intimidating to exclude
many species, while the hilltop must be distinctive and
visible at large distances. Some 33 families of Diptera
are known to hilltop (Skevington 2008).
188 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
OVERVIEW OF TAXA
Lower Diptera
Some of the most common and easily recognized flies
(Figs. 9.1A-E, 9.3C), including black flies (Simuliidae),
crane flies (Tipuloidea), fungus gnats (e.g., Mycetophil-
idae and Sciaridae), and mosquitoes (Culicidae), belong
to the lower Diptera (also known as the ‘Nematocera’).
The group contains approximately 40 families and
more than 52,000 species worldwide (Evenhuis et al.
2007). Although the Diptera and several subordinate
taxa (e.g., Brachycera, Eremoneura, Cyclorrhapha, and
Schizophora) are considered monophyletic, the lower
Diptera generally are considered a paraphyletic or
grade-level grouping (Hennig 1973, Wood and Borkent
1989, Oosterbroek and Courtney 1995, Yeates and
Wiegmann 1999, Yeates et al. 2007). Despite this posi-
tion, a review is useful of some of the features shared
by members of this phyletic grade of Diptera. For the
most part, adults of lower Diptera are characterized
as slender, delicate, long-legged flies with long, mul-
tisegmented antennae (e.g., Culicidae, Tanyderidae,
and Tipulidae); however, the group also includes some
stout-bodied flies with relatively short antennae (e.g.,
Axymyiidae, Scatopsidae, and Simuliidae). Larvae of
most lower Diptera have a well-developed, sclerotized
head capsule (Figs. 9.2A-F) (Courtney et al. 2000).
Although a few lineages in the lower Diptera (e.g.,
Bibionomorpha) occur primarily in terrestrial or
semiterrestrial habitats, the vast majority of lower
Diptera have larvae and pupae that are aquatic or
semi-aquatic (Foote 1987, Brown 2001, Merritt et al.
2003). Aquatic habitats include a wide range of lentic
(standing water) and lotic (flowing water) situations
(Courtney and Merritt 2008, Courtney et al. 2008).
Lakes, cold and hot springs, temporary pools, stagnant
waters of ground pools, phytotelmata (tree holes and
other plant cavities), and artificial containers (e.g.,
buckets and tires) are among the many lentic habitats
colonized by larvae. The Culicoidea (e.g., Chaoboridae,
Culicidae, and Dixidae) are especially well represented
in lentic habitats. These families include proficient
swimmers that can travel to considerable depths;
yet their larvae generally remain near the water
surface because of the dependence on atmospheric
respiration. The culicid genera Coquillettidia and
Mansonia are unusual because their larvae use their
specialized respiratory siphons to obtain oxygen
from submerged or floating vegetation (Wood et al.
1979, Clements 1992). Free-swimming larvae of
common midges (Chironomidae) and biting midges
(Ceratopogonidae) do not depend on atmospheric
respiration and can colonize larger and deeper bodies
of water. Some chironomids survive at great depths,
with one species, Sergentia koschowi, known to occur as
deep as 1360 m in Lake Baikal (Linevich 1971). Lotic
habitats of lower Diptera range from slow, silty rivers
to torrential streams to groundwater zones (Courtney
and Merritt 2008). The larvae of the net-winged
midges (Figs. 9.2F, 9.3E) (Blephariceridae), mountain
midges (Fig. 9.2D) (Deuterophlebiidae), and black
flies (Simuliidae) are among the most specialized
inhabitants of flowing waters; all lack spiracles
(exchanging oxygen directly through their cuticle)
and have structural modifications that permit survival
on current-exposed substrates. Blepharicerid larvae,
which frequently occur in current velocities exceeding
2 m/sec, show perhaps the greatest morphological
specialization, including ventral suctorial discs used
to adhere to smooth rocks (Zwick 1977; Hogue
1981; Courtney 2000a, 2000b). Similar habitats and
comparably unusual attachment devices (prolegs
with apical rows of hooks) are typical of larval
Deuterophlebiidae (Courtney 1990, 1991) and
Simuliidae (Crosskey 1990, Adler et al. 2004). Other
specialized lotic habitats include seepages on cliff faces
and waterfall splash zones, where larval Thaumaleidae
and many Chironomidae, Psychodidae, Simuliidae,
and Tipulidae can be common (Vaillant 1956, 1961,
Sinclair and Marshall 1987, Sinclair 1988, 1989,
2000, Craig and Currie 1999), and saturated wood
along stream margins, where larvae of Axymyiidae
and certain Tipuloidea (e.g., Lipsothrix) reside (Dudley
and Anderson 1987, Wood 1981). Groundwater zones
are another important but largely unstudied habitat
for larval Diptera, particularly for the Chironomidae
(McElravy and Resh 1991, Ward 1992, 1994).
Finally, larvae of a few lower Diptera (e.g., some
members of the Ceratopogonidae, Chironomidae,
Culicidae, and Tipulidae) can be abundant in marine
and brackish-water environments, including intertidal
pools, seaweed beds, lagoons, and estuarine marshes
(Hashimoto 1976, Linley 1976, O’Meara 1976,
Robles and Cubit 1981, Pritchard 1983, Colbo 1996,
Cranston and Dimitriadis 2005, Dimitriadis and
Cranston 2007).
The taxonomic and ecological diversity of the lower
Diptera is reflected in the wide range of larval feeding
habits, which encompass nearly every trophic group.
Biodiversity of Diptera 189
Many groups consume live plants (e.g., Cecidomyiidae
and some Tipuloidea) or decomposing plant fragments
or fungi (e.g., Mycetophilidae, Sciaridae, and many
Tipuloidea). Others feed on decaying, fine organic
matter and associated microorganisms (e.g., many
Chironomidae). The larvae of some aquatic families
(e.g., Blephariceridae and Thaumaleidae) use special-
ized mouthparts to graze on the thin film of algae and
organic matter on rocks and other substrates (Courtney
2000a, 2000b; Alverson et al. 2001). Many families
contain a few predaceous species, whereas the larvae
of some groups (e.g., Ceratopogonidae) feed primar-
ily or exclusively on other animals (McAlpine et al.
1981, H¨ovemeyer 2000). Nearly all of these trophic
groups are represented in the diverse family Chironomi-
dae (nearly 7000 species) and superfamily Tipuloidea
(more than 15,000 species). Their trophic diversity
and numerical abundance make the lower Diptera an
important component in aquatic and terrestrial ecosys-
tems, both as primary consumers and as a food resource
for other invertebrates, fish, amphibians, reptiles, birds,
and mammals. The Chironomidae, which in aquatic
ecosystems are often the most abundant organisms in
both numbers and biomass, can be especially important
in ecosystem functioning (Armitage et al. 1995). The
trophic importance of aquatic Diptera extends also to
aquaculture programs in which nearly every life stage
can be an important component of fish diets.
Brachycera
Lower Brachycera
As in the lower Diptera, the lower Brachycera
are paraphyletic, but remain a convenient grade
for discussion. This group is also widely known
as the ‘Orthorrhapha’ (referring to the T-shaped
opening of the pupal exuviae) and comprises mostly
predaceous larvae (except Stratiomyomorpha) and
parasitoids of spiders and other insect orders. Adult
lower Brachycera are blood feeders, predators, or
flower visitors. The lower Brachycera contain some
of the largest and most colorful flies, including bee
flies, horse flies, mydas flies, and robber flies. This
grade includes some 24,000 species comprising
20 families assigned to three infraorders (Stratiomy-
omorpha, Tabanomorpha, and Xylophagomorpha)
and several superfamilies (Asiloidea and Nemestri-
noidea) (Yeates et al. 2007).
The Pantophthalmidae are enormous flies (up
to 5.5 cm in length), with larvae that dig galleries
in dead or living trees and likely feeding on the
fermenting sap in the tunnels (Val 1992, D. M. Wood,
personal communication). Both the Xylomyidae and
Stratiomyidae (Fig. 9.1N) are unique among the
lower Brachycera in regard to their scavenging and
filter-feeding habits (Rozkoˇsn ´y 1997) and by pupating
in the final-instar larval exuviae (comparable to the
cyclorrhaphan puparium). The Stratiomyidae larvae
can be assigned generally to two groups: terrestrial
and aquatic. Terrestrial larvae live in decaying
leaves and other plant material, upper layers of soil,
manure, under loose bark of decaying trees, and in
ant nests. Aquatic larvae (Fig. 9.2L) can be found
in saturated moss, littoral zones of ponds, lakes, and
marshes, hygropetric situations in spring streams,
phytotelmata, roadcuts or similar seepages, saline
habitats, and even hot thermal springs (Rozkoˇsn´y
1997, Sinclair and Marshall 1987, Sinclair 1989).
Feeding on vertebrate blood by female flies has
evolved at least two or three times in the lower Brachyc-
era, but is restricted to the Tabanomorpha (Athericidae,
Rhagionidae sensu lato, and Tabanidae) (Wiegmann
et al. 2000, Grimaldi and Engel 2005). The Tabanidae
(deer flies and horse flies) are well known to campers
and swimmers during the early summer months in
northern latitudes due to the voracious blood-sucking
behavior of most species. Many species of the subfamily
Pangoniinae are characterized by their long mouth-
parts (known as long-tongues), often stretching longer
than their than body length. These groups generally are
believed to be nectar feeders (Goldblatt and Manning
2000), but several species also have been observed
feeding on warm-blooded (humans – Philoliche;
Morita 2007) and cold-blooded (caimans – Fidena;
B.A. Huber, personal communication) vertebrates.
Tabanid larvae mostly inhabit swampy biotopes, where
they prey on insect larvae. They even are known to
feed opportunistically on toads (Jackman et al. 1983).
Adults of many of the remaining families of
lower Brachycera are fast-flying flower visitors.
Moegistorhynchus longirostris (Nemestrinidae) from
southern Africa possesses a proboscis nearly five
times its body length and is an important pollinator
of tubular flowers (Goldblatt and Manning 2000).
Bee flies (Bombyliidae) occur worldwide and reach
their greatest diversity in Mediterranean climates
(Yeates 1994). The female abdomen of several bee
fly subfamilies is modified to form an invaginated
190 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
(a)
(c)
(f)
(j)
(m) (n) (o)
(k) (l)
(g) (i)
(h)
(d) (e)
(b)
Fig. 9.1 Adult Diptera. (a) Tipulidae
(Tanysipetra) habitus, dorsal view.
(b) Axymyiidae (Axymyia), lateral view.
(c) Limoniidae (Prionolabis)mating
pair, oblique-dorsal view. (d) Bibionidae
(Bibio) habitus, oblique-lateral view.
(e) Culicidae (Culex) feeding on ranid
frog. (f) Empididae (Empis) habitus,
lateral view. (g) Pipunculidae taking
flight, oblique-lateral view.
(h) Micropezidae (Grallipeza) habitus,
lateral view. (i) Diopsidae (Teleopsis)
head, frontal view. (j) Conopidae
(Stylogaster) mating pair, lateral view.
(k) Asilidae (Proctacanthus) feeding on
dragonfly, oblique-dorsal view.
(l) Sarcophagidae (Sarcophaga) habitus,
dorsal view. (m) Scathophagidae
(Scathophaga) habitus, oblique-lateral
view. (n) Stratiomyidae habitus, lateral
view. (o) Calliphoridae (Hemipyrellia)
habitus, frontolateral view. (See color
plate). (Images by E. Bernard [a], G.
Courtney [b, c, h, i, m], S. Marshall [e, f,
g, j, k], M. Rice [d] and I. Sivec [l, n, o].)
sand chamber, which is first filled when they alight
on open surfaces (Yeates 1994, Greathead and
Evenhuis 1997). The egg is laid in the sand chamber
and coated with soil particles before being ejected
by the hovering female onto oviposition sites. The
larvae of the Bombyliidae are mostly parasitoids
of holometabolous insects (e.g., acridoid egg pods,
solitary bees, and wasps) (Greathead and Evenhuis
1997). The larvae of small-headed flies (Acroceridae)
are internal parasitoids of true spiders. First-stage
larvae actively seek out hosts, capable of looping along
a single web strand (Nartshuk 1997). In contrast to
adults of most lower Brachycera, those of the Asilidae
(Fig. 9.1K) are strictly predaceous on insects (Hull
1962). They focus on large prey from a wide variety of
insect orders, sometimes taking prey more than twice
their size (e.g., dragonflies; Platt and Harrison 1995).
Empidoidea
The dance flies, balloon flies, and other predaceous flies
(Fig. 9.1F) that traditionally have been placed in the
Biodiversity of Diptera 191
Fig. 9.2 Larval Diptera.
(a) Tipulidae (Epiphragma) habitus,
dorsal (top) and ventral (bottom)
views. (b) Ptychopteridae
(Bittacomorpha) head, thorax and
abdominal segments I–III, lateral
view. (c) Nymphomyiidae
(Nymphomyia) habitus lateral view.
(d) Deuterophlebiidae (Deuterophlebia)
habitus, dorsal view. (e) Psychodidae
(Pericoma) habitus, lateral view.
(f) Blephariceridae (Horaia) habitus,
dorsal (left) and ventral (right) views.
(g) Calliphoridae (Lucilia) habitus,
dorsal view. (h) Tephritidae (Eurosta)
habitus, ventral view. (i) Syrphidae
(Syrphus) feeding on aphids, dorsal
view. (j) Syrphidae (Microdon)on
glass, lateral view. (k) Sciomyzidae
(Tetanocera) habitus, lateral view.
(l) Stratiomyidae (Caloparyphus)
habitus, dorsal view. (See color plate).
(Images by G. Courtney [a–f, h, k, j]
and S. Marshall [g, i, l].)
(a) (b)
(d)
(c)
(e)
(f) (g)
(i) (j)
(k) (l)
(h)
family Empididae are now classified in four families
of Empidoidea, along with the long-legged flies of the
family Dolichopodidae. With approximately 12,000
described species and many more undescribed species,
the Empidoidea are one of the largest superfamilies
of Diptera and the most diverse lineage of predaceous
flies (Sinclair and Cumming 2006). The vast majority
are predators as adults, with the few exceptions being
obligate flower-feeding groups that consume pollen as
their only protein source. They are found in a vari-
ety of forested and open habitats where they breed in
moist soils, decaying wood, and dung, and occur in
aquatic habitats. All known larvae appear to be preda-
tors on invertebrates. The common name ‘dance flies’
is derived from the behavior of members of the large
subfamily Empidinae in which adult males transfer
nuptial gifts to the female during courtship and mating
(Cumming 1994). The Dolichopodidae are common
metallic-colored flies, often observed sitting on leaves
and mud flats. Many possess elaborate leg ornamen-
tations that are used in courtship displays (Sivinski
1997).
192 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
(a) (b)
(d)(c) (e)
(f) (g)
(i) (j)
(h)
Fig. 9.3 Scanning electron
micrographs of Diptera.
(a) Phoridae (Thaumatoxena) adult
habitus, lateral view.
(b) Nymphomyiidae (Nymphomyia)
adult head, lateral view.
(c) Blephariceridae (Blepharicera)
adult head, frontal view.
(d) Phoridae (Termitophilomya)
adult head, lateral view.
(e) Blephariceridae (Agathon) larva
habitus, oblique-frontal view.
(f) Ptychopteridae (Bittacomorpha)
larval mouthparts, ventral view.
(g) Athericidae (Atherix) larval
head, lateral view.
(h) Calliphoridae (Onesia) larval
head, ventral view.
(i) Sarcophagidae (Metopia) larval
head, oblique-ventral view.
(j) Calliphoridae (Bellardia) larval
head, ventral view. (Images by G.
Courtney [b, c, e, f, g] and T. Pape
[a, d, h–j].)
Lower Cyclorrhapha
This group of taxa has been considered a monophyletic
group (Aschiza) by some workers (McAlpine 1989,
Disney 1994a). As with the lower Diptera, strong
evidence now suggests that the lower Cyclorrhapha
are a paraphyletic assemblage (Cumming et al. 1995,
Sinclair and Cumming 2006, Moulton and Wiegmann
2007). The seven lower cyclorrhaphan families are
discussed below.
The Platypezidae include 250 species, commonly
found individually hovering in deep shade or in
swarms of dancing males in forest openings. Males of
Microsania (‘smoke flies’) form epigamic swarms in
smoke from forest and campfires. Eggs of platypezids
are laid between the gills or in the pores of fungi
on which the larvae feed. The Ironomyiidae occur
in both dry sclerophyll and rainforest habitats and
have been found hilltopping (Skevington 2008). The
family contains one described and two undescribed
extant species (all in the genus Ironomyia), all from
Australia (D. K. McAlpine, personal communication).
Fifteen fossil species from the Holarctic Region
have been described from five additional genera
from the Upper Jurassic and Cretaceous periods
(McAlpine 1973, Zhang 1987, Mostovski 1995,
Biodiversity of Diptera 193
Grimaldi and Cumming 1999). The Lonchopteridae
are distinctive, pointed-winged, strongly bristled,
often yellowish-brown flies, most commonly found in
moist environments, along streams or ponds, in bogs,
in deciduous forests, or even in alpine meadows or
hot springs (Nielsen et al. 1954, Smith 1969). Some
species are found in hot, dry meadows (Andersson
1970), while others occur in rocky tidal zones near the
coast (Dahl 1960). Most species are bisexual but at least
one species, Lonchoptera bifurcata, is parthenogenetic
in most areas of its nearly cosmopolitan range (Stalker
1956). It is the only species of the family that is found
in the neotropics and Australia. Lonchopterids have a
predominantly Old World distribution, with 46 of the
approximately 60 described species occurring in the
Palearctic and Oriental Regions. Larvae are apparently
saprophagous, microphagous, or mycetophagous, but
more study is needed to confirm these feeding habits.
The Phoridae, or scuttle flies, are one of the most
diverse fly families and have been proposed as the most
biologically diverse family of insects on Earth (Disney
1994b). Approximately 4000 described species have
been described but more than 30,000 species are esti-
mated to exist (Evenhuis et al. 2007, Brown 2008).
More than half of the described phorid species belong
to the huge genus Megaselia. They are found in almost
all terrestrial habitats, with the exclusion of exception-
ally cold and dry environments. Larvae have extremely
diverse tastes, with many saprophagous species, fun-
givores, and herbivores (including leaf miners, root
feeders, and bud, seed, and fruit feeders). The majority
of phorid larvae are likely predators, parasitoids, or
parasites, and several species occur in highly special-
ized habitats such as pitcher plants and even intertidal
areas (Disney 1998).
The Pipunculidae, or big-headed flies, comprise
1400 described species. Most pipunculids that have
been studied are endoparasitoids of several families
of Homoptera (Auchenorrhyncha). The only known
exceptions are species of the genus Nephrocerus,which
attack adult crane flies (Tipuloidea) (Koenig and
Young 2007). Adults use all terrestrial habitats, but
diversity and numbers are greatest in forest openings
and along forest edges. Pipunculids also are known for
their hilltopping behavior (Skevington 2000, 2001).
Adult flower flies (Syrphidae), also known as hover
flies, vary considerably in size and appearance, ranging
from 4 to 25 mm long and from small black flies to
large wasp or bee mimics. They include about 6000
described species and are among the most abundant
and conspicuous flies. Their visibility is partly related to
their ability to hover motionless and partly to their fre-
quent flower visitation. They are among the most signif-
icant pollinators in the Diptera and should be assessed
in comparison to the bees (Ssymank et al. 2008). Some
species such as Episyrphus balteatus are strong fliers
and are migratory. Many larvae are entomophagous,
feeding on ant brood, aphids and other soft-bodied
Sternorrhyncha, and social wasp larvae (Thompson
and Rotheray 1998). Saprophagous syrphid larvae
exploit wet or moist conditions and are typically asso-
ciated with fermenting tree sap, rot holes and fallen
wood, decaying vegetation in water or wet compost,
and dung. Some larvae are mycophages– phytophages,
feeding in pockets of decay in live plants.
Non-Calyptratae Muscomorpha
The acalyptrates are likely another grade, or group of
convenience. Treated as the sister group of the calyp-
trates for years (Hennig 1971, 1973; McAlpine 1989),
some evidence suggests that this assemblage is para-
phyletic (Griffiths 1972). The early radiation of the
more than 80 families of flies in this group appears
to have been explosive, which may have obscured the
evolutionary history of the group. Griffiths (1972),
Hennig (1973), and McAlpine (1989) are the only
researchers to have proposed a phylogeny (or at least
a phylogenetically based classification) for the entire
group. Despite their efforts, evidence indicates that
even some of the superfamilies are not monophyletic.
This radiation resulted in a remarkable diversity of
flies and many of the groups are of considerable
importance to society. More than 50% of the aca-
lyptrate species are contained in just six large families:
Agromyzidae, Chloropidae, Drosophilidae, Ephydridae,
Lauxaniidae, and Tephritidae. Some of the ecological
diversity shown by these families and others is discussed
below.
Most of the Conopidae are parasitoids of bees and
aculeate Hymenoptera. One lineage, the Stylogas-
trinae, comprises parasitoids of orthopteroid insects.
Many of the Neotropical species follow foraging army
ant raids, attacking the orthopteroids that flee from the
advancing army. All known Psilidae are phytophagous
and some are well-known pests (e.g., the carrot rust
fly, Chamaepsila hennigi; Peacock and Norton 1990,
as Psila rosae) (Szwejda and Wrzodak 2007). Some
Diopsidae are agricultural pests on grasses, but some
also are well-known research subjects because of
194 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
their variously developed eyestalks. Many species are
sexually dimorphic for the length of the eyestalks,
with males having much longer eyestalks than
females. This dimorphism is believed to have evolved
because of the mating advantages that they bestow
on these males. Females show a strong preference
for males with longer eyestalks, and males compete
with each other to control lekking aggregation sites
by a ritualized contest that involves facing each other
and comparing their relative eyespans, often with the
front legs spread out to add emphasis. A few other
families of acalyptrate flies have members that have
developed similar types of apparent runaway sexual
selection. For example, some Drosophilidae, Platy-
stomatidae, Richardiidae, and Tephritidae have stalked
eyes, or large, sometimes antler-like spines extending
from their genae. Of these families, the Platystomatidae
and Tephritidae are important phytophagous groups.
Tephritids in particular include many pest species that
are becoming problems because they are spread via
global trade (e.g., Oriental fruit fly, Bactrocera orientalis,
and Mediterranean fruit fly, Ceratitis capitata). The
Drosophilidae are most famous because of Drosophila
melanogaster, the subject of considerable genetics
research. Drosophilids also are renowned for their
expansive radiation in Hawaii, where approximately
1000 species (more than 500 described) radiated from a
single colonizer (Kaneshiro 1997, O’Grady et al. 2003).
The Pyrgotidae are bizarre-looking flies, most com-
monly seen at lights at night. These flies are specialized
internal parasitoids of scarab beetles and likely have
a pronounced effect in controlling some populations
of pest scarabaeids (Steyskal 1987). The Piophilidae
are small flies that tend to specialize on carrion. One
species, the cheese skipper Piophila casei, is a serious pest
in the food industry, with larvae found in cured meats,
smoked fish, cheeses, and decaying animals. Some
species, such as Protopiophila litigata, form impressive
mating aggregations on discarded cervid antlers (Bon-
duriansky and Brooks 1998). The Clusiidae are another
of a handful of acalyptrate families known to engage
in this type of lekking behavior. Males establish dom-
inance at a lekking site by defending territories from
other males on logs or branches to attract females and
mate (Lonsdale and Marshall 2004).
Larvae of all Agromyzidae feed on living plant
tissues, forming mines that are species specific. Most are
either monophagous or oligophagous, and although
best known as leaf miners, they attack all parts of
plants (Dempewolf 2004).
Sciomyzid larvae are all predators or parasitoids of
freshwater or terrestrial molluscs (Berg and Knutson
1978). The Chamaemyiidae are free-living predators
of adelgids, aphids, coccids, and scales and have been
used in biological control programs (Gaimari and
Turner 1996, Vail et al. 2001). The Sphaeroceridae are
diverse and associated with all types of organic decay
including dung, carrion, fungi, supralittoral seaweed,
compost, mammal nests, conifer duff, cave debris, and
deposits of dead vegetation (Marshall and Richards
1987). The Ephydridae, or shore flies, are important
food for wildlife along both freshwater and saltwater
pools. In some wetlands, such as Mono Lake in Cali-
fornia, millions of birds are supported almost entirely
by ephydrids (Jehl 1986, Rubega and Inouye 1994).
Larvae of the Chloropidae have varied food habits
(Ferrar 1987). Many are phytophagous, damaging
cereals and other grasses. These species include the frit
fly (Oscinella frit), the wheat stem maggot (Meromyza
americana), and the gout fly of wheat and barley in
Europe (Chlorops pumilionis). Others are saprophagous,
fungivorous, and even predaceous. Thaumatomyia
glabra is an important predator on the sugarbeet root
aphid Pemphigus populivenae. One of the most unusual
habits among Chloropidae is that of the species of
the Australian genus Batrachomyia, whose larvae live
under the skin on the back of frogs (Sabrosky 1987).
Calyptratae
The calyptrate flies are generally rather robust, most
are strong fliers, and many are in the size range
of the common housefly. The group contains some
22,000 species, which are arranged in 10–15 fam-
ilies, depending on the classification (Evenhuis et al.
2007). A large number of species breed in living or
decaying plant or fungal material (especially the mus-
coid families Anthomyiidae, Fanniidae, Muscidae, and
Scathophagidae). The biology of species in the large
family Muscidae is particularly varied, with habitats
including vertebrate dung and carrion; organic debris
in nests, burrows, and dens of mammals, birds, and
insects; rot holes and decaying wood; sap runs; liv-
ing plants; fungi; and in or on the edges of ponds
and streams (Skidmore 1985, Ferrar 1987). Evolution
can take surprising routes, as in the Scathophagidae,
whereby a few lineages have evolved from the ancestral
plant-feeding life habit into breeding in dung or rotting
seaweed and even as predators on caddisfly egg masses
or small invertebrates (Kutty et al. 2007).
Biodiversity of Diptera 195
Large mammals are a rich source of food for many
calyptrates, which may suck their blood, imbibe their
sweat, eat their dung, or even be true endoparasites. The
tse-tse (Glossinidae) and the ectoparasitic louse and bat
flies (Hippoboscidae, including the Nycteribiinae and
Streblinae) have excelled as highly specialized blood
feeders and, like the few other calyptrate bloodsuckers,
both males and females take a blood meal. Hosts
are mainly mammals and birds but a few feed on
large reptiles. The reproductive biology of hippoboscoid
calyptrates is remarkable in that the eggs hatch one at
a time in the female oviduct, and the larva is nourished
by a secretion from the female accessory glands until it
is fully fed and near pupariation (Ross 1961, Hill 1963,
Marshall 1970, Potts 1973). Obligate parasites of mam-
mals are found particularly in the family Oestridae,
with larvae taking up their final position either sub-
dermally (Cuterebrinae, Hypodermatinae, and lower
Gasterophilinae), in the gastrointestinal tract (higher
Gasterophilinae), or in the nasopharyngeal cavities
(Oestrinae) (Zumpt 1965).
Numerous calyptrates are associated with either
vertebrate or invertebrate carrion, and some species
infest wounds or body orifices of living vertebrates as
larvae. Social insects are hosts of many calyptrates.
Several blow flies (Calliphoridae) are associated
with ants and termites as either kleptoparasites
or, more rarely, parasitoids (Ferrar 1987, Sze et al.
2008). Solitary, aculeate Hymenoptera can have
their nests usurped by kleptoparasitizing flesh flies
(Sarcophagidae) of the subfamily Miltogramminae
(Spofford and Kurczewski 1990, Pape 1996). Snails
and earthworms are heavily exploited by both flesh flies
and blow flies (Keilin 1919, Ferrar 1976, Guimar˜aes
1977, Downes 1986), and pterygote insects are hosts
to the exclusively parasitoid species of the large family
Tachinidae (Stireman et al. 2006).
SOCIETAL IMPORTANCE
As expected for a ubiquitous group with diverse habits
and habitats, the Diptera are of considerable economic
importance. Pestiferous groups can have significant
effects on agriculture, animal and human health, and
forestry. Other groups can be a general nuisance when
present in large numbers or because of allergic reactions
to detached body setae. Despite these negative effects,
flies play a valuable role as scavengers, parasitoids
and predators of other insects, pollinators, food for
predators, bioindicators of water quality, and tools for
scientific research.
Diptera as plant pests (agriculture,
silviculture, and floriculture)
A large number of fruit flies (Tephritidae) are capable
of causing considerable economic damage to fruits
and vegetables, making these flies perhaps the most
important dipteran family to agriculture (e.g., Dowell
and Wange 1986, McPheron and Steck 1996, Norrbom
2004). The genera Anastrepha, Bactrocera, Ceratitis,
Dacus,andRhagoletis contain most of the pest species.
Economic impact includes direct losses from decreasing
yield, increasing costs for control and fruit treatment,
and shrinking export markets due to local regulations.
Quarantine laws designed to reduce the spread of fruit
fly species can severely restrict global commerce of
many commercial fruits. Evidence of their economic
effects is illustrated by the millions of US dollars spent
annually to prevent the Mediterranean fruit fly from
entering California (Jackson and Lee 1985, Dowell and
Wange 1986, Aluja and Norrbom 1999).
The Agromyzidae, well known for the plant-mining
habits of their larvae, also contain a number of
important plant pests, including the chrysanthe-
mum leafminer (Liriomyza trifolii), the serpentine
leafminer (L. brassicae), and the vegetable leafminer
(L. sativae) (Spencer 1973, 1990). The pea leafminer
L. huidobrensis is a highly polyphagous species that
can damage a wide range of field and greenhouse
crops, including alfalfa, artichoke, beans, beets,
carrots, celery, lettuce, melons, onions, peas, potato,
pumpkin, spinach, tomatoes, and several crucifers and
ornamentals. Control of established populations can
be especially problematic (Steck 2005).
Among lower Diptera, the gall midges (Cecidomyi-
idae) are perhaps the best-known agricultural pests.
A widespread and common group containing mostly
plant-feeding species, gall midges are suspected to
include several thousand new and undescribed
species; however, studies of the tropical fauna are
still in their infancy (e.g., Gagn´e 1994). Because of
the feeding habits of their larvae, gall midges include
many serious plant pests, especially species that attack
cereal crops and conifers. As with many fruit flies and
other plant pests, their effects include not only direct
damage, but also economic losses related to quarantine
issues (Pollard 2000, Gagn´e et al. 2000).
196 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
A few Diptera can be important floricultural pests.
Larvae of the black fungus gnats (Sciaridae) Bradysia
coprophila and Bradysia impatiens feed on roots and algae
in the upper soil surface, which can cause considerable
damage in propagation areas and seedling flats. Larval
feeding also can cause wilting and facilitate entry of
plant pathogens. Furthermore, adult flies are known
to disseminate soil-inhabiting pathogens on their bod-
ies and in their feces (Parrella 2004). The ephydrid
Scatella stagnalis can be a pest in ornamental nurs-
eries and greenhouses, where adult flies spread fungal
spores. Fecal spots left on leaves by resting adults
also can cause cosmetic damage to plants (Parrella
2004).
Several Phoridae and Sciaridae and the moth fly
Psychoda phalaenoides can be important pests of com-
mercial mushroom gardens (Hussey 1960, Rinker
and Snetsinger 1984, Somchoudhury et al. 1988,
Scheepmaker et al. 1997, Menzel and Mohrig 1999).
Although larval feeding can cause moderate damage,
the major effect is through the transmission of fungal
diseases (White 1981).
The Bombyliidae are usually considered beneficial
insects because some species parasitize and prey on
cutworms, beetle grubs, and grasshopper egg pods;
however, Heterostylum robustum kills up to 90% of the
larvae of the alkali bee, an important alfalfa pollina-
tor in northwestern USA (Bohart et al. 1960, Bohart
1972). Other pests in apiaries include the European
miltogrammine flesh fly Senotainia tricuspis,which
infects as many as 90% of the adult bees in a hive
(Santini 1995a, 1995b, Palmeri et al. 2003). Honey-
bees are sometimes attacked by species of the phorid
Melaloncha (Ram´
ırez 1984, Brown 2004, Gonzalez and
Brown 2004), and native colonies of meliponine bees
can be affected by the phorids Pseudohypocera kerteszi
and Megaselia scalaris (Robinson 1981; Reyes 1983;
Hern´andez and Guti´errez 2001; Robroek et al. 2003a,
2003b). Megaselia scalaris also can cause consider-
able damage to live arthropod cultures in laboratories,
insect zoos, and butterfly houses (Disney 1994b; G. W.
Courtney, personal observations).
Medical and veterinary importance
Disease transmission
The mouthparts of many adult Diptera have effective
piercing stylets, enabling these flies to ‘bite’ and
suck blood. Major families with piercing and sucking
mouthparts include the bat flies, biting midges,
black flies, horse and deer flies, louse and bat flies,
mosquitoes, phlebotomine sand flies (Psychodidae),
tse-tse, and a few muscid flies. Because of their
blood-feeding habits, these flies are natural carriers of
pathogens and play a major role in the transmission
of bacteria, fungi, nematodes, protozoans, viruses,
and other parasites. The affinities of some dipterans to
carrion and excrement might enhance their capacity
to transmit disease agents, and for this reason alone
Diptera can be considered the most economically
important insect order. Diptera-borne diseases affect
humans, as well as livestock worldwide, and the
resulting costs are enormous.
Mosquitoes are perhaps the best-known and
most-studied blood-feeding dipterans, due largely
to their medical and veterinary importance. Of the
approximately 3600 known species of mosquitoes,
fewer than 150 are pests or vectors of pathogens that
cause disease in humans and domesticated animals
(Harbach 2007). However, these species, which are
largely confined to the genera Aedes (traditional, broad
sense), Anopheles,andCulex, are the indirect cause
of more morbidity and mortality among humans
than any other group of organisms. Mosquitoes are
vectors of a number of agents that cause debilitating
diseases, including malaria, yellow fever, filariasis,
dengue, dog heartworm, the encephalitides, and
related viral diseases. For malaria alone, the effects
are staggering: 300–500 million people are infected
annually; 1.0–1.5 million people die every year (World
Health Organization 2007); an African child dies from
malaria every 30 sec (WHO 2007); 35 million future
life-years are lost because of premature mortality and
disability (World Bank 1993); and an annual cost
of nearly US$2 billion is incurred in tropical Africa
(MicrobiologyBytes 2007). Even in contemporary
North America, the effect of mosquito-borne diseases
remains acute. A recent (2002–2003) outbreak
of West Nile virus in Louisiana came at a price of
approximately US$20 million, with slightly more
than half the costs related to the illness (e.g., direct
medical costs and productivity losses from illness and
death) and the remaining costs related to public health
responses (e.g., mosquito control, surveillance, and
abatement) (KPLC 2004).
Black flies transmit the filarial parasites Dirofilaria (in
bears) and Mansonella and Onchocerca (in humans), as
well as the protozoans Leucocytozoon (in birds) (Crosskey
Biodiversity of Diptera 197
1990). Phlebotomine sand flies are vectors of the agents
that cause leishmaniasis (sand fly fever) and also can
transmit filarial parasites (Icosella neglecta) to the edible
European green frog (Desportes 1942). Biting midges
are capable of transmitting at least 66 viral pathogens
and a wide range of microorganisms (Borkent 2004),
including those responsible for livestock diseases that
cause blue-tongue in sheep and cattle, African horse
sickness, bovine ephemeral fever, and eastern equine
encephalitis (Parsonson 1993, Mellor and Boorman
1995, Wall and Shearer 1997). The Tabanidae serve
as vectors of African eye-worm or loa loa that causes
loiasis, the bacterium that causes tularaemia, and the
Old World trypanosome Trypanosoma evansi (Oldroyd
1973).
Calyptrate flies (e.g., house, stable, and blow
flies) harbor more than 100 species of pathogenic
microorganisms (Greenberg 1971, 1973, F¨orster et al.
2007, Sawabe et al. 2006). Many species of Musca
are carriers of bovine and equine filariases, as well as
bacteria and viruses. For example, Musca autumnalis
can transmit Parafilaria bovicola, and eyeworms of
the genus Thelazia,andHydrotaea irritans serves as a
mechanical vector of Corynebacterium pyogenes,acause
of mastitis in cattle (Neville 1985, Krafsur and Moon
1997). Musca sorbens transmits eye diseases such as
trachoma and conjunctivitis (Emerson et al. 2000).
The horn fly Haematobia irritans is a well-established
biting cattle pest throughout many tropical and
temperate areas of the Northern Hemisphere, while
its close relative, the buffalo fly Haematobia exigua is
particularly important to cattle and dairy industries of
Australia.
As important vectors of blood-borne diseases,
dipterans have shaped human culture (Harrison
1978). This role is evident especially in Africa,
where trypanosome-infected tse-tse can be a serious
constraint on livestock practices, and epidemics of
sleeping sickness have had profound socioeconomic
implications (e.g., Lyons 1992, Hide 1999). In the
New World, the introduction of yellow fever has had a
comparable impact (Crosby 2006). Between 1904 and
1914, the need to control yellow fever and malaria was
a major contributor to the completion of the Panama
Canal (Powers and Cope 2000). During military
campaigns, diseases such as malaria can account
for more casualties than the fighting (Bruce-Chwatt
1988), which explains the interest in vector-borne
disease research at military-affiliated institutions (e.g.,
Walter Reed Army Institute of Research, Walter
Reed Biosystematics Unit). The original transmission
of the HIV virus from chimpanzees to humans is
suspected to have been caused by stable flies or similar
blood-sucking flies (Eigen et al. 2002), adding yet
another example of the vast influence of Diptera on
human societies. Mosquitoes have even shaped human
evolution through their disease-carrying capabilities,
with the most notable example from Africa, where
the sickle-cell anemia gene became prevalent due
to its partial protection against the malaria parasite
(Pagnier et al. 1984, Barnes 2005).
Myiasis
The families Calliphoridae (blow flies), Oestridae (bot
flies), and Sarcophagidae (flesh flies) are the major pro-
ducers of myiasis, a term referring to the development
of dipteran larvae in a living vertebrate body. Bot
flies are involved in dermal, enteric, and nasopharyn-
geal myiasis of animals and sometimes humans. The
larvae of cattle grubs migrate through the host’s body
and eventually reach the upper back where they cut
a small opening in the hide and remain there until
ready to pupate. Economic losses result from reduction
in milk production, weight loss, and damage to hides
(Scholl 1993). In the Northern Hemisphere, Hypoderma
bovis and H. lineatum (Hypodermatinae) are the major
pests, whereas in the New World tropics, Dermatobia
hominis (Cuterebrinae) is the prevalent cattle warble
fly (Guimar˜aes et al. 1983). Production losses can be
significant, with annual losses in Brazil estimated at
US$200–260 million (Grisi et al. 2002). Most bot flies
either never attack humans, or do so only accidentally
(Zumpt 1965). However, Dermatobia hominis,known
also as the human bot fly or t´orsalo, develops read-
ily in humans. Infections are painful but generally
benign, even when the larva is allowed to develop to
maturity (Dunn 1930). In rare cases, infections can be
lethal (Rossi and Zucoloto 1973, Noutis and Millikan
1994). Damage caused by nasal bot flies (Oestrinae)
in camels, goats, and sheep, and stomach bot flies
(Gasterophilinae) in donkeys and horses varies from
violent reactions (i.e., ‘gadding’ behavior) caused by
the ovipositing flies, to irritation by larvae when bur-
rowing into oral tissues and subsequent interference
with digestion. These attacks can reduce growth rates
and are particularly harmful to younger individuals
(Zumpt 1965).
Myiasis-producing blow flies and flesh flies usually
are attracted to the wounds and sores of humans and
198 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
domestic animals, where larvae feed on necrotic tissue
and accidentally can be ingested or invade wounds,
causing severe discomfort and subsequent secondary
infections. Certain calliphorids can cause severe
primary myiasis, particularly Cochliomyia hominivorax
(the primary or New World screwworm) and Lucilia
cuprina and Chrysomya bezziana in the Old World tropics
(Hall and Wall 1995). In recent years, other species
of screwworms in the genus Chrysomya have been
introduced accidentally from the Old World into South
America and have spread north into Central America,
two even reaching North America (Baumgartner
1993, Tomberlin et al. 2001). Among the flesh flies,
species of Wohlfahrtia cause myiasis in commercially
raised mink (Eschle and DeFoliart 1965), livestock
(Hall 1997, Valentin et al. 1997, Farkas and Kepes
2001, Farkas et al. 2001) and, rarely, humans (Hall
and Wall 1995, Delir et al. 1999, Iori et al. 1999).
Human urogenital myiasis occasionally is caused by
larvae of the Psychodidae and Phoridae (Disney and
Kurahashi 1978, Abul-Hab and Salman 1999). Species
of the African bot fly genus Gedoelstia may larviposit in
the eyes of cattle, goats, and sheep, with the encephalitic
form often fatal (Zumpt 1965). Cases of human oph-
thalmomyiasis have been caused by Gedoelstia spp.,
Hypoderma spp., and Oestrus ovis (Bisley 1972, Masoodi
and Hosseini 2004, Lagac´e-Wiens et al. 2008).
Invasive alien Diptera
The introduction of alien species can have devastat-
ing effects on the native fauna. The effects have been
documented for the avifauna of Hawaii, with the intro-
duction of avian malaria and a suitable vector, Culex
quinquefasciatus (Van Riper et al. 2002). In contrast, the
avifauna of the Gal´apagos Islands is largely intact, but
the establishment of this avian vector and potential dis-
eases poses a great threat (Whiteman et al. 2005). The
introduction of the nestling parasite Philornis downsi to
the Gal´apagos Islands also represents a serious threat
to the endemic passerine fauna (Fessl et al. 2006).
Numerous dipterans have become invasive around
the world, especially members of the Agromyzidae,
Anthomyiidae, Calliphoridae, Culicidae, Drosophili-
dae, Muscidae, Phoridae, Psilidae, Sarcophagidae, and
Tephritidae, but our attention to the phenomenon is
strongly skewed toward pests and disease vectors. The
most detailed studies of the effect of invasive dipterans
on local species have thus been for the blow flies (Wells
1991) and mosquitoes (Juliano and Lounibos 2005).
Diptera as a general nuisance
In addition to their significance in myiasis and the trans-
mission of disease agents, flies can be a general nuisance
and interfere with human activities (e.g., Cook et al.
1999, Howard 2001). The nuisance problem includes
harassment by mosquitoes, gnats, and other flies, and
the occasional presence of Diptera in true nuisance
numbers (e.g., Westwood 1852). Several modern-
day examples of the latter exist. Following European
settlement of Australia, the Australian bush fly, Musca
vetustissima, bred vigorously in the cow dung that accu-
mulated in the absence of native ruminant-adapted
dung beetles. South African dung beetles imported
during the 1960s–1980s have ameliorated the prob-
lem (Ridsdill-Smith 1981, Matthiessen et al. 1984,
Ridsdill-Smith et al. 1987). Another example, from
Europe, pertains to the chloropid Thaumatomyia notata.
This fly, in attempting to find suitable overwinter-
ing sites, enters apartments by the millions (Nartshuk
2000, Kotrba 2004, Nartshuk and Pakalniˇskis 2004).
Swarms of this species have even been mistaken as
smoke, prompting calls to the fire brigade (Kiesenwet-
ter 1857, Letzner 1873). Species from several other
families (e.g., Phoridae and Sphaeroceridae) prolifer-
ate indoors such that eradication measures are needed
(Fredeen and Taylor 1964; Disney 1991, 1994b; Cle-
worth et al. 1996). Certain Psychodidae can be a
nuisance as both larvae and adults, the former when
mass occurrences in trickling filters of sewage treat-
ment facilities diminish flow and filtering efficiency,
and adults when setae from the wings and body are
inhaled, causing a disease similar to bronchial asthma
(Gold et al. 1985).
Adults of the aquatic families Chaoboridae and
Chironomidae sometimes constitute a nuisance
by their sheer numbers emerging from ponds
and lakes. When encountering swarms of these
flies, avoiding inhaling them or keeping them
out of one’s eyes can be difficult. Emerging
chironomid adults can be a serious nuisance of
lakefront settlements and cities (Ali 1991), and
massive swarms can cause traffic problems (Lindegaard
and J´onasson 1979). At some East African lakes,
swarms of emerging Chaoborus edulis are so dense as
to pose a risk of suffocation should one get trapped
within them, particularly if swarms also contain the
allergenic chironomid midge Cladotanytarsus lewisi
(Armitage et al. 1995). Despite these negative effects,
local people capture the swarming midges and convert
Biodiversity of Diptera 199
them into round cakes that are dried in the sun for
later consumption (Oldroyd 1964, Eibl and Copeland
2005).
Even beneficial species of flies, such as Sarcophaga
aldrichi (‘the friendly fly’), build up to such large
numbers that people complain vehemently about their
presence. This species is an important parasitoid of the
forest tent caterpillar (Malacosoma disstria), completely
controlling their massive outbreaks (USDA, Forest Ser-
vice 1985).
Biting midges, black flies, deer flies and horse flies,
and mosquitoes, apart from their capacity as disease
vectors, are infamous for their sometimes incessant
harassment of humans and other animals. In some
parts of the world, considerable resources are devoted
to reducing the numbers of mosquitoes and black flies
along major rivers (e.g., Skovmand 2004) and in cities
(e.g., Callaway 2007). Mosquitoes, as well as biting
midges, deer flies, and horse flies, also can be abun-
dant along beaches and in salt marshes and mangrove
swamps. The latter are especially well-known breeding
sites for species of Culicoides, abundances of which can
manifest in biting rates of several hundred per minute
(Borkent 2004).
In many urban areas, the major nuisance fly is the
common housefly Musca domestica, which often occurs
in large numbers around humans or human activ-
ities. In addition to the discomfort caused by direct
contact with large numbers of flies buzzing around
food, garbage, and other items, house flies can be
mechanical vectors of various microbial pathogens
(Nayduch et al. 2002, Sanchez-Arroyo 2007).
Other dipterans that annoy and interfere with
human comfort include certain members of the
Chloropidae, especially the genus Hippelates,com-
monly known as eye gnats. Larvae of these flies inhabit
the soil, but adults can be a nuisance because they
are attracted to sweat, tears, and other secretions
around the eyes or exposed skin. Though more of an
annoyance than a health risk, eye gnats can serve
as vectors of the agents of anaplasmosis, bacterial
conjunctivitis, and bovine mastitis (Lindsay and
Scudder 1956, Mulla 1965, Tondella et al. 1994).
Diptera in biological control
A large number of Diptera can be beneficial, espe-
cially the many predaceous or parasitoid groups that
help regulate insect pest populations. Many species
are native components of the ecosystems in which
they are found, but others are introduced to con-
trol native or exotic pests. The gall midge Feltiella
acarisuga is a widespread and effective predator of
spider mites (Tetranychidae) (Gagn´e 1995). It can
successfully control populations of Tetranychus urticae
in various crops (Opit et al. 1997) and is a poten-
tially useful agent for integrated pest management
of spider mites in greenhouses (Gillespie et al. 1998).
The aphid predatory midge Aphidoletes aphidimyza is
another gall midge that has shown promise as an effec-
tive predator of aphids. It is an important component
of biological control programs for greenhouse crops
and is now widely available commercially (Hoffmann
and Frodsham 1993). The cottony cushion scale killer
Cryptochetum iceryae (Cryptochetidae), along with the
vedalia beetle Rodolia cardinalis, was introduced from
Australia to California to successfully control the cot-
tony cushion scale Icerya purchasi (DeBach and Rosen
1991, Waterhouse and Sands 2001). Florida citrus
growers subsequently introduced C. iceryae for the same
purpose, and the species is now widespread in warmer
parts of the New World (Pitkin 1989). Another group
used in the control of coccids and aphids is the fam-
ily Chamaemyiidae. Leucopis tapiae, for example, was
introduced into Hawaii to control the Eurasian pine
adelgid (Greathead 1995, Vail et al. 2001).
Snail-killing flies (Sciomyzidae) control populations
of the intermediate hosts of trematodes causing
bilharzia (Berg and Knutson 1978, Maharaj et al.
1992) and multivoltine species have the potential
to control pest helicid snails in Australian pastures
(Coupland and Baker 1995). Some studies (Graham
et al. 2003, Porter et al. 2004) suggest that
ant-decapitating species of Phoridae can suppress
introduced populations of fire ants. The Pipunculidae
are of interest as potential control agents for rice and
sugarcane leafhopper pests (Greathead 1983). Species
of the muscid genus Coenosia, whose larvae and adults
are predaceous, are being tested for biological control
of agromyzids, ephydrids, sciarids, and white flies in
greenhouses (K ¨uhne 2000).
The exclusively parasitic Tachinidae are used exten-
sively in biological control programs, especially against
pestiferous Lepidoptera. Success stories include the
introduction of Bessa remota to control coconut moths
in Fiji (DeBach and Rosen 1991) and the use of Bil-
laea claripalpis, Lixophaga diatraeae,andLydella minense
to control sugarcane stem borers (Diatraea spp.) in the
Neotropical Region (Bennett 1969, Cock 1985, DeBach
200 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
and Rosen 1991). Other introductions have achieved
limited success, such as use of the Palearctic tachinid
Cyzenis albicans to control the winter moth Operoph-
tera brumata in Canadian deciduous forests (Horgan
et al. 1999). Although most tachinids attack a nar-
row spectrum of hosts, a few species (e.g., Compsilura
concinnata) have been reared from hundreds of different
host species. Under the latter circumstances, a potential
negative side effect of biological control is the harm to
nontarget hosts. For example, C. concinnata was intro-
duced to control gypsy moths in New England. The
tachinid had a modes effect on the target host but is
thought to have led to declining populations of local
silk moths (Boettner et al. 2000).
A few Diptera, particularly fruit flies (Tephritidae),
have been used for biological control of weeds. Primary
targets have been knapweeds and thistles (Carduus,
Centaurea,andCirsium) and various other genera with
noxious species (e.g., Ageratina, Lantana,andSenecio)
(Bess and Haramoto 1972, White and Clement 1987,
Harris 1989, White and Elson-Harris 1992, Turner
1996). Other flies used for the control of certain weeds
include a few Agromyzidae (Spencer 1973), Syrphidae,
and Cecidomyiidae (Gagn´e et al. 2004).
Pollination
Diptera are major contributors to the maintenance of
plant diversity through their participation in many pol-
lination systems and networks (Ssymank et al. 2008).
Diptera probably were among the first angiosperm
pollinators, and flies might have been influential
in spurring the early angiosperm diversification
(Labandeira 1998, Endress 2001). Flies visit flowers to
obtain nectar for energy and pollen for protein; flowers
also provide species-specific rendezvous sites for
mating and a beneficial microclimate (Kearns 2002,
Kevan 2002). Diptera are among the most common
insects that visit flowers (Free 1993), and in Belgium
more than 700 plant species in 94 families were
visited by flower flies alone (De Buck 1990). Diptera
have lower flower-visiting consistency, are generally
much less hairy than the aculeate Hymenoptera, and
most lack specialized structures for pollen transport.
Despite the latter, pollen clings to flies with furry body
vestiture and some flies have foretarsal modifications
allowing them to gather and eat pollen (e.g., Holloway
1976, Neff et al. 2003). But even generalist flower
visitors contribute significantly to plant reproductive
success (Kearns 2001, Kevan 2002). Diptera pollinate
a significant number of important crops, including
apples, cacao, carrots, cashew, cassava, cauliflower,
leek, mango, mustard, onions, strawberries, and tea
(Heath 1982, Hansen 1983, Clement et al. 2007,
Mitra and Banerjee 2007). The highly specialized
flowers of cacao are pollinated exclusively by small
midges (Ceratopogonidae), particularly of the genus
Forcipomyia (Young 1986, 1994), and an increasing
number of flowering plants are being discovered that
depend entirely on dipteran pollinators. Examples
include the seed-for-seed mutualism where species of
the anthomyiid genus Chiastocheta pollinate the closed
flowers of Trollius europaeus (Pellmyr 1989), and the
gall-midge pollination of Artocarpus, which is a mutu-
alism involving also a parasitic fungus (Sakai et al.
2000). A significant number of flowers have specialized
in being pollinated by carrion flies, including the
world’s largest flower Rafflesia arnoldii (Beaman et al.
1988). Diptera are particularly important pollinators
of flat to bowl-shaped flowers in habitats and under
conditions where bees are less active. Many flies have
adapted well to moist and cool habitats, such as cloud
forests and arctic and alpine environments, which have
a large proportion of dipteran pollinators (Kevan 1972;
Elberling and Olesen 1999; Kearns 1992, 2001). Small
Diptera might be the most important pollinators in
the forest understory, particularly for shrubs with
numerous small, inconspicuous, and dioecious flowers
(Larson et al. 2001, Borkent and Harder 2007).
Other ecological services (scavengers
and decomposers)
In most terrestrial and freshwater ecosystems, Diptera
are more species rich and have a higher biomass than
do other insect decomposers (McLean 2000). Represen-
tatives of many families, including the Calliphoridae,
Coleopidae, Muscidae, Mycetophilidae, Phoridae, Psy-
chodidae, Sarcophagidae, Sciaridae, Sepsidae, Sphae-
roceridae, Stratiomyidae, Syrphidae, and Tipuloidea
are important decomposers and recyclers of decaying
organic matter. The importance of Diptera in recy-
cling dung has been well studied (Laurence 1953,
1954, 1955; Papp 1976, 1985; Papp and Garz ´o 1985;
Skidmore 1991; O’Hara et al. 1999).
The black soldier fly Hermetia illucens (Stratiomyi-
dae), a pantropical species that breeds in decaying
fruit and other decomposing organic material (James
Biodiversity of Diptera 201
1935), exemplifies the decomposing capacity and
ecological significance of flies. Chicken manure
colonized by H. illucens often leads to reduced amounts
of manure (Sheppard et al. 1994), results in fewer
houseflies (Sheppard 1983, Axtell and Arends 1990),
and can even provide a food resource (i.e., prepupae)
for fish and swine (Newton et al. 1977, Bondari
and Sheppard 1981). Hermetia illucens also has been
implicated as a potentially beneficial species to the
citrus industry, where the destruction of orange-peel
waste can be a costly endeavor. Pape (2009) recounted
a compelling example of this potential use in Costa
Rica, where waste from a local orange-growing
company dumped hundreds of tons of orange peel on
strongly degraded bushland during the dry season.
At the onset of the rains, populations of H. illucens
boomed and the waste was completely decomposed
by the larvae in nine months. As an added value, a
new indigenous dry forest was resprouting from this
‘Biodiversity Processing Ground’.
As effective decomposers of organic material, many
Diptera have a high pest potential through their ability
to locate and infect stored human food. Provisions par-
ticularly prone to become infested are household meats
and meat products, which may become ‘blown’ with
blow fly eggs. Cheese and ham also are favored habitats
for the cheese skipper Piophila casei. An important way
of preserving fish practically throughout the world is to
dry the meat under the sun, and a number of flies breed
in such cured fish, especially blow flies of the genera
Calliphora, Chrysomya,andLucilia. These flies can be
a serious problem in many tropical and subtropical
societies, causing losses of up to 30% (Haines and Rees
1989, Esser 1991, Wall et al. 2001).
Diptera of forensic, medicolegal,
and medical importance
Flies are usually the first insects to arrive at vertebrate
carrion, which make especially the Calliphoridae (and
species of Fanniidae, Muscidae, Phoridae, Piophilidae,
Sarcophagidae, and Stratiomyidae) potential forensic
indicators in cases involving dead bodies. A forensic
entomologist estimates the time elapsed since a blow
fly larva, found on a corpse, hatched from its egg
by backtracking the development time, that is, by
measuring the number of degree days required to
complete development and subtracting this from the
known total required for complete larval development
(Higley and Haskell 2001), or through phenological
information (Staerkeby 2001).
The affinity of blow fly maggots for decaying flesh
makes some species ideal for cleaning certain wounds,
particularly bedsores and age- and diabetes-related
gangrene involving reduced circulation. Even severe
burns and extensive abrasions, where small islands of
dead tissue scattered over larger areas make physical
removal complicated, can be treated in this manner.
The use of maggot therapy for treating wounds of
humans and livestock is an old discovery (Grantham-
Hill 1933, Leclercq 1990, Sherman et al. 2000), but,
since it has been improved through sterile breeding
of larvae and controlled application under specially
developed bandages, the technique has been taken up
by many clinics. The maggots provide a dual effect by
eating the dead tissue and secreting antiseptic saliva.
Even the mechanical stimulus from the active larvae
exerts a micromassage promoting the circulation of
lymphatic fluids in the recovering tissues (Sherman
2001, 2002, 2003).
Diptera as research tools
Physiology and genetics
The muscle tissues of vertebrates and insects might
be only remotely homologous (Mounier et al. 1992),
but insect indirect flight muscles bear some functional
and physiological resemblances to human heart mus-
culature (Chan and Dickinson 1996, Maughan et al.
1998). Knowledge of functional properties and organi-
zation of proteins in Drosophila wing-muscle myofibrils
(e.g., Vigoreaux 2001) and the expression of heterol-
ogous human cytoplasmic actin in Drosophila flight
muscles carry significant potential for increased under-
standing of human muscular disorders (Brault et al.
1999). Asynchronous flight muscles, whereby a sin-
gle nerve impulse causes a muscle fiber to contract
multiple times, is the key to the extreme mechanical
and physiological efficiency behind the high-frequency
wing beat necessary for sustained flight in many insects.
Peak performance is found in some Ceratopogonidae in
which Sotavalta (1947, 1953) measured frequencies
surpassing 1000 Hertz in Forcipomyia sp., and by exper-
imentally reducing wing length, more than doubled
this frequency. This can be accomplished only when
muscles are able to contract in an oscillatory man-
ner, requiring that they are attached to an appropriate
202 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
mechanically resonant load, which in a fly would be
its thorax and wings. Insect flight muscles, therefore,
offer insights into muscular operational design, con-
tractile costs, and energy-saving mechanisms (Conley
and Lindstedt 2002, Syme and Josephson 2002), which
will have implications for human health as well as for
biotechnological advances.
Drosophila melanogaster was introduced as a
laboratory animal for the geneticist about a century
ago (Castle 1906). This introduction turned out to be
extremely fruitful, and D. melanogaster is now, for many
people, the icon of genetic research. Numerous studies
using this model species have brought tremendous
insight into gene expression, gene regulatory mech-
anisms, and, more recently, genomics (Ashburner
and Bergman 2005). Revealing the D. melanogaster
genome, which was the second animal genome to
be fully sequenced, has been remarkably rewarding.
Almost 75% of the candidate human disease genes
can be matched by homologues in Drosophila (Reiter
et al. 2001), and today FlyBase (http://flybase.bio
.indiana.edu/), the Drosophila community database,
is providing one of the highest-quality annotated
genome sequences for any organism. The recognition
of gene homologues carries a large potential for
improved treatments of human disorders ranging from
type-II diabetes to alcoholism (e.g., Campbell et al.
1997, Fortini et al. 2000, Brogiolo et al. 2001, Leevers
2001, Morozova et al. 2006). At a higher genetic
level, the discovery of the Hox genes in D. melanogaster
in the early 1980s paved the way for entirely new
insights into how organisms regulate the identity of
particular segments and body regions by controlling
the patterning along the embryonic head-to-tail axis
(e.g., Carroll 1995, Akam 1998, Lewis 1998). The
short generation time of D. melanogaster, which makes
it so suitable for genetic studies, also makes it suitable
for studies on age-specific and lifetime behavior
patterns involved in aging (Carey et al. 2006), as
well as on the genetics and physiology of age-related
memory impairment (Horiuchi and Saitoe 2005).
Technology
Insect flight uses thin, flexible plates (wings) reinforced
by a system of ridges (the veins) that allow for
semi-automated deformation, which optimizes aerody-
namic forces. Insect wings typically produce two–three
times more lift than can be accounted for by conven-
tional aerodynamics, and they produce a high amount
of lift while keeping drag at a minimum. These features
make insects attractive models for microplane design
(Dwortzan 1997, Ellington 1999, Wooton 2000,
Bar-Cohen 2005). Dipteran flight has been fine-tuned
through millions of years of evolution, and some of
the most diligent insect flyers are found among the
Diptera, whereby certain species of bee flies, flower
flies, pipunculids, and rhiniine blow flies show a range
of aerial acrobatics unsurpassed by any other flying
animal. Such maneuverability built into so-called
micro-air-vehicles (MAVs), apart from obvious
military interests, would be of use, for example, in
aerial surveillance operations and reconnaissance in
confined spaces, for example, by rescue squads dealing
with partly collapsed or burning buildings.
Female mosquitoes have excelled for millions of
years in their ability to take a blood meal almost
without being felt by the victim. As they cut their
way through tough vertebrate skin, the apically
microserrated mandibular and maxillary stylets
provide less friction and require lower insertion
forces, compared with conventional human-made
syringe needles. Micro-engineers and biomedical
engineers, therefore, have turned to the mosquito
proboscis in an attempt to develop ultra-narrow
syringe needles for minimally invasive (pain-free),
micro-electromechanical drug delivery and sampling
of body fluids for microdialysis in continued medical
monitoring (Cohen 2002, Gattiker et al. 2005).
Diptera in conservation
Bioindicators
Biomonitoring is the use of living organisms or
their responses to evaluate environmental quality
(Rosenberg and Resh 1993, Resh et al. 1996, Barbour
et al. 1999, Moulton et al. 2000, Bonada et al. 2006,
Rosenberg et al. 2007) and involves three general
areas of investigation: (1) surveys before and after an
impact to determine the effects of that impact (e.g.,
Thomson et al. 2005), (2) regular sampling or toxicity
testing to measure compliance with legally mandated
environmental quality standards (e.g., Yoder and
Rankin 1998, Maret et al. 2003), and (3) large-scale
surveys to establish reference conditions or evaluate
biological impairment across geographical landscapes
and under different land-management practices
(e.g., Klemm et al. 2003, Black and Munn 2004).
Biodiversity of Diptera 203
Although these studies encompass a range of taxa and
assessment metrics, many will include in their analyses
some evaluation of Diptera diversity and abundance.
The inclusion of Diptera has been especially true of
water-quality and bioassessment studies to classify the
degree of pollution or other impacts in a water body.
The mouthparts of many aquatic larvae continuously
filter detritus and microorganisms, and the associated
habitat and microhabitat specialization means that
species associations can be informative about water
quality (e.g., Sæther 1979). The Chironomidae are
perhaps the most widely used dipterans for these
purposes. Larvae of the midge genus Chironomus are
commonly referred to as ‘blood worms’ because of
hemoglobin in their blood, a trait that permits survival
in poorly oxygenated aquatic habitats. These larvae
and those of other Chironomidae (e.g., Wiederholm
1980, Raddum and Saether 1981) and other aquatic
Diptera (e.g., the moth flies Psychoda and rat-tailed
maggots Eristalis) are often used as indicators of
polluted water or water low in oxygen (e.g., Lenat
1993b, Barbour et al. 1999, Courtney et al. 2008).
Furthermore, evaluation of morphological deformities
in the larval head capsules (e.g., changes in sclerite
shape) of dipterans, especially chironomids, has
been used extensively to assess environmental stress
(Wiederholm 1984; Warwick and Tisdale 1988;
Warwick 1989, 1991; Diggins and Stewart 1993;
Lenat 1993a). Although the general perception is that
most aquatic Diptera are tolerant of environmental
impacts, some groups (e.g., Blephariceridae and
Deuterophlebiidae) have received tolerance values
indicating extreme sensitivity to environmental
perturbations (Lenat 1993b, Barbour et al. 1999,
Courtney et al. 2008). Despite these general patterns,
most families have species that exhibit a range of
tolerances from pristine to impaired conditions.
The Chironomidae, Culicidae, and Tipuloidea are
noteworthy in this respect. Finally, some Diptera
(especially Chironomidae) are used commonly in
acute and chronic laboratory toxicity studies to
compare toxicants and the factors affecting toxicity,
and ultimately to predict the environmental effects of
the toxicant (Michailova et al. 1998, Karouna-Renier
and Zehr 1999).
Diptera are rich in species with specific microhabi-
tat or breeding-site requirements, providing them with
a high potential for habitat-quality assessment and
conservation planning (Rotheray et al. 2001). Abun-
dance patterns of stiletto flies (Therevidae) can be an
indicator of habitat heterogeneity and successional
stage in dry areas (Holston 2005). Haslett (1988)
used flower flies as bioindicators of environmental
stress on ski slopes in Austria, and Sommagio (1999)
suggested that their strength would be in the evaluation
of landscape diversity. Many saproxylic and fungivo-
rous flies have an association with old-growth forests,
which implies a considerable potential as indicators
of woodland quality that might help in designing and
implementing management strategies such as forest-
cutting regimes and tree-species composition (Speight
1986; Økland 1994, 1996, 2000; Good and Speight
1996; Fast and Wheeler 2004; Økland et al. 2004,
2008). Special measures have been taken to conserve
saproxylic insects in England (Rotheray and Mac-
Gowan 2000). Increasing forest cover and changes in
their management in the Netherlands since the 1950s
have meant that the saproxylic Syrphidae in general
are on the increase (Reemer 2005). A similar situa-
tion has been indicated for Germany (Ssymank and
Doczkal 1998). The ecologically diverse Dolichopodi-
dae are showing promise as indicators of site value
over a range of nonforest habitats (Pollet 1992, 2001;
Pollet and Grootaert 1996), and endemic Hawaiian
Dolichopodidae, together with selected Canacidae, Chi-
ronomidae, and Ephydridae, are potential indicators
of valuable aquatic habitats with high native diversity
(Englund et al. 2007).
Vanishing species
The world biota is under ever-increasing pressure from
humankind. Only four dipterans are on the IUCN Red
List of Threatened Species (IUCN 2007), containing
species that are either vulnerable, endangered, or crit-
ically endangered, and therefore facing a higher risk
of global extinction, but this list is probably grimly
misleading, as still more species of Diptera are finding
their way onto regional red lists (e.g., Falk 1991, Stark
1996, Binot et al. 1998, Pollet 2000). Geographically
restricted populations are often particularly vulnera-
ble, and numerous species of Diptera most probably
have already disappeared due to the arrival of invasive
species on oceanic islands or through major habitat
destruction. An example is the single flesh fly endemic
for Bermuda, Microcerella bermuda, which has not been
collected for the last 100 years and was not recovered
in the most recent inventory (Woodley and Hilburn
1994). Other Diptera are so rare that they face immi-
nent extinction: the peculiar Mormotomyia hirsuta,sole
204 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
representative of the family Mormotomyiidae, has been
found in only a single bat roost in a large, cave-like
rock crevice in Kenya (Oldroyd 1964, Pont 1980). The
three species of rhino stomach bot flies (Gyrostigma spp.)
have been experiencing increasing difficulties main-
taining healthy populations, with gradually declining
host stocks. The situation already could be critical for
Gyrostigma sumatrensis, which is still known only from
larvae expelled from captive Sumatran rhinos in a few
European zoos, all before 1950. The African G. con-
jungens, known from the black rhino, has not been
captured since 1961. A successful conservation pro-
gram for the white rhino, however, appears to have had
positive effects on G. rhinocerontis (Barraclough 2006).
The Delhi Sands flower-loving fly, Rhaphiomidas termi-
natus abdominalis (Mydidae), is the first fly to be listed
as endangered by the US Endangered Species Act. It is
endemic to the Delhi Sands formation, a small area of
ancient inland dunes in southern California, where the
adults are nectar feeders and the drastic loss of habitat
has led to its perceived decline and endangered status
(Rogers and Mattoni 1993).
What may be the first dipteran to be eradicated
by humans is the European bone skipper Thyreophora
cynophila. When described by Panzer (1794), this fly
was rather common and was often observed in Aus-
tria, France, and Germany. A beautiful, redheaded fly, it
could be observed walking on big cadavers such as dead
dogs, horses, and mules in the early spring (Robineau-
Desvoidy 1830). Suddenly, 50 years after its discovery,
it disappeared and has never been collected again. Its
disappearance might be due to changes in livestock
management and improved carrion disposal, following
the Industrial Revolution in Europe, but the underlying
scenario probably is the reduction of the megafauna,
including the near absence of large predators to leave
large carcasses with partly crushed long bones, thereby
limiting access to the medullar canal and bone marrow,
the favored breeding site for T. cynophila. The Quater-
nary megafauna extinctions, which might have had
a human component, most probably had fatal conse-
quences for those Diptera that we assume depended on
these large animals or their excreta. The stomach bot
fly (Cobboldia russanovi) of the woolly mammoth disap-
peared with its host (Grunin 1973), and D. K. McAlpine
(2007) envisions a much larger fauna of Australian
wombat flies at the time of the large marsupials some
100,000 years ago. A few other species have been
declared globally extinct: four species of Emperoptera
and one species of Campsicnemus (Dolichopodidae), and
a single species of Drosophila (Drosophilidae), all from
Hawaii (Hawaii Biological Survey 2002, IUCN 2007),
and the volutine Stoneyian tabanid (Stonemyia velutina)
from California (IUCN 2007).
Diptera as part of our cultural legacy
Just as much as man might have realized ‘that his
destiny is coupled to coexistence with a complex biota
that also contains Diptera’ (Pape 2009), flies are an
integral part of our cultural past. Thousands of chil-
dren have been fascinated by the fairy tale about the
brave little tailor, who got ‘seven in one blow’, and who
has not been laughing at jokes where a customer in a
restaurant complains that ‘there is a fly in my soup!’
Some might even have read ‘The Fly’ by George Lan-
gelaan (1957), featuring a human–housefly hybrid, or
seen one or more of the several films based on this short
story, or any of the sequels to these. Flies were one of the
biblical plagues (Exodus 8:21), but fly symbolism spans
the entire range from William Golding’s (1954) somber
novel ‘Lord of the Flies’ to William Blake’s (1794) lyric
poem ‘The Fly’:
Little Fly
Thy summers play,
My thoughtless hand
Has brush’d away.
Am not I
A fly like thee?
Or art not thou
A man like me?
For I dance
Anddrinkandsing;
Till some blind hand
Shall brush my wing.
If thought is life
And strength and
breath;
And the want
Of thought is death;
Then am I
A happy fly,
If I live,
Or if I die.
Table 9.1 Families of Diptera and numbers of described species in the world. Family classification and species richness based on Evenhuis et al. (2007).
Superfamily Described
Suborder Infraorder Other Category (or equivalent) Family Species
‘LOWER DIPTERA’ Ptychopteromorpha Ptychopteridae 74
‘LOWER DIPTERA’ Ptychopteromorpha Tanyderidae 38
‘LOWER DIPTERA’ Culicomorpha Chironomoidea Ceratopogonidae 5621
‘LOWER DIPTERA’ Culicomorpha Chironomoidea Chironomidae 6951
‘LOWER DIPTERA’ Culicomorpha Chironomoidea Simuliidae 2080
‘LOWER DIPTERA’ Culicomorpha Chironomoidea Thaumaleidae 173
‘LOWER DIPTERA’ Culicomorpha Culicoidea Chaoboridae 55
‘LOWER DIPTERA’ Culicomorpha Culicoidea Corethrellidae 97
‘LOWER DIPTERA’ Culicomorpha Culicoidea Culicidae 3616
‘LOWER DIPTERA’ Culicomorpha Culicoidea Dixidae 185
‘LOWER DIPTERA’ Blephariceromorpha Blephariceroidea Blephariceridae 322
‘LOWER DIPTERA’ Blephariceromorpha Blephariceroidea Deuterophlebiidae 14
‘LOWER DIPTERA’ Blephariceromorpha Nymphomyioidea Nymphomyiidae 7
‘LOWER DIPTERA’ Bibionomorpha Axymyioidea Axymyiidae 6
‘LOWER DIPTERA’ Bibionomorpha Bibionoidea Bibionidae 754
‘LOWER DIPTERA’ Bibionomorpha Bibionoidea Hesperinidae 6
‘LOWER DIPTERA’ Bibionomorpha Bibionoidea Pachyneuridae 5
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Bolitophilidae 59
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Cecidomyiidae 6051
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Diadocidiidae 19
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Ditomyiidae 93
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Keroplatidae 907
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Lygistorrhinidae 30
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Mycetophilidae 4105
‘LOWER DIPTERA’ Bibionomorpha Sciaroidae Rangomaramidae 39
‘LOWER DIPTERA’ Bibionomorpha Sciaroidea Sciaridae 2224
‘LOWER DIPTERA’ Psychodomorpha Anisopodidae 158
‘LOWER DIPTERA’ Psychodomorpha Perissommatidae 5
‘LOWER DIPTERA’ Psychodomorpha Psychodidae 2886
‘LOWER DIPTERA’ Psychodomorpha Scatopsoidea Canthyloscelidae 16
‘LOWER DIPTERA’ Psychodomorpha Scatopsoidea Scatopsidae 323
‘LOWER DIPTERA’ Psychodomorpha Scatopsoidea Valeseguyidae 3
‘LOWER DIPTERA’ Tipulomorpha Tipuloidea Cylindrotomidae 67
‘LOWER DIPTERA’ Tipulomorpha Tipuloidea Limoniidae 10,334
‘LOWER DIPTERA’ Tipulomorpha Tipuloidea Pediciidae 494
(continued)
205
Table 9.1 (continued).
Superfamily Described
Suborder Infraorder Other Category (or equivalent) Family Species
‘LOWER DIPTERA’ Tipulomorpha Tipuloidea Tipulidae 4325
‘LOWER DIPTERA’ Tipulomorpha Trichoceridae 160
Subtotal (‘Lower
Diptera’)
52,302
BRACHYCERA Stratiomyiomorpha Pantophthalmidae 20
BRACHYCERA Stratiomyiomorpha Stratiomyidae 2666
BRACHYCERA Stratiomyiomorpha Xylomyidae 134
BRACHYCERA Tabanomorpha Athericidae 122
BRACHYCERA Tabanomorpha Austroleptidae 8
BRACHYCERA Tabanomorpha Oreoleptidae 1
BRACHYCERA Tabanomorpha Rhagionidae 707
BRACHYCERA Tabanomorpha Spaniidae 43
BRACHYCERA Tabanomorpha Tabanidae 4387
BRACHYCERA Xylophagomorpha Xylophagidae 136
BRACHYCERA Vermileonomorpha Vermileonidae 59
BRACHYCERA Muscomorpha ‘lower Brachycera’ Nemestrinidae 275
BRACHYCERA Muscomorpha ‘lower Brachycera’ Acroceridae 394
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Apioceridae 169
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Apsilocephalidae 3
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Apystomyiidae 1
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Asilidae 7413
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Bombyliidae 5030
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Evocoidae 1
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Hilarimorphidae 32
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Mythicomyiidae 346
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Mydidae 463
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Scenopinidae 414
BRACHYCERA Muscomorpha ‘lower Brachycera’ Asiloidea Therevidae 1125
BRACHYCERA Muscomorpha Empidoidea Atelestidae 10
BRACHYCERA Muscomorpha Empidoidea Brachystomatidae 145
BRACHYCERA Muscomorpha Empidoidea Dolichopodidae 7118
BRACHYCERA Muscomorpha Empidoidea Empididae 2935
BRACHYCERA Muscomorpha Empidoidea Homalocnemus group 7
BRACHYCERA Muscomorpha Empidoidea Hybotidae 1882
BRACHYCERA Muscomorpha Empidoidea Iteaphila group 27
BRACHYCERA Muscomorpha Empidoidea Oreogeton group 12
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Ironomyiidae 1
206
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Lonchopteridae 58
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Opetidae 5
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Phoridae 4022
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Pipunculidae 1381
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Platypezidae 252
BRACHYCERA Muscomorpha ‘lower Cyclorrhapha’ Syrphidae 5935
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Nerioidea Cypselosomatidae 34
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Nerioidea Megamerinidae 15
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Nerioidea Micropezidae 578
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Nerioidea Neriidae 111
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Diopsidae 183
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Gobryidae 5
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Nothybidae 8
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Psilidae 321
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Somatiidae 7
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Strongylophthalmyiidae 47
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Syringogastridae 10
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Diopsoidea Tanypezidae 21
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Conopoidea Conopidae 783
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Ctenostylidae 10
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Lonchaeidae 480
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Pallopteridae 66
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Piophilidae 82
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Platystomatidae 1162
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Pyrgotidae 351
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Richardiidae 174
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Tachiniscidae 3
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Tephritidae 4621
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Tephritoidea Ulidiidae 672
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Lauxanioidea Celyphidae 116
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Lauxanioidea Chamaemyiidae 349
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Lauxanioidea Eurychoromyiidae 1
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Lauxanioidea Lauxaniidae 1893
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Coelopidae 35
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Dryomyzidae 25
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Helcomyzidae 12
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Helosciomyzidae 23
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Heterocheilidae 2
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Huttonidae 8
(continued)
207
Table 9.1 (continued).
Superfamily Described
Suborder Infraorder Other Category (or equivalent) Family Species
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Natalimyzidae 1
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Phaeomyiidae 3
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Rhopalomeridae 33
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Sciomyzidae 604
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sciomyzoidea Sepsidae 375
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Agromyzidae 3013
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Anthomyzidae 94
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Asteiidae 132
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Aulacigastridae 18
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Clusiidae 349
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Fergusoninidae 29
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Marginidae 3
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Neminidae 14
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Neurochaetidae 20
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Odiniidae 62
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Opomyzidae 61
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Periscelididae 84
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Teratomyzidae 8
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Opomyzoidea Xenasteiidae 13
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Acartophthalmidae 4
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Australimyzidae 9
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Braulidae 7
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Canacidae 119
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Carnidae 90
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Chloropidae 2863
208
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Cryptochetidae 33
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Inbiomyiidae 10
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Milichiidae 276
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Carnoidea Tethinidae 193
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sphaeroceroidea Chyromyidae 106
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sphaeroceroidea Heleomyzidae 717
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sphaeroceroidea Mormotomyiidae 1
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sphaeroceroidea Nannodastiidae 5
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Sphaeroceroidea Sphaeroceridae 1580
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Ephydroidea Camillidae 40
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Ephydroidea Curtonotidae 61
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Ephydroidea Diastatidae 48
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Ephydroidea Drosophilidae 3925
BRACHYCERA Muscomorpha Schizophora: ‘Acalyptrates’ Ephydroidea Ephydridae 1977
BRACHYCERA Muscomorpha Schizophora: Calyptratae Hippoboscoidea Glossinidae 23
BRACHYCERA Muscomorpha Schizophora: Calyptratae Hippoboscoidea Hippoboscidae 786
BRACHYCERA Muscomorpha Schizophora: Calyptratae Muscoidea Anthomyiidae 1896
BRACHYCERA Muscomorpha Schizophora: Calyptratae Muscoidea Fanniidae 319
BRACHYCERA Muscomorpha Schizophora: Calyptratae Muscoidea Muscidae 5153
BRACHYCERA Muscomorpha Schizophora: Calyptratae Muscoidea Scathophagidae 392
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Calliphoridae 1524
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Rhiniidae 363
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Mystacinobiidae 1
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Oestridae 150
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Rhinophoridae 147
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Sarcophagidae 2632
BRACHYCERA Muscomorpha Schizophora: Calyptratae Oestroidea Tachinidae 9629
Subtotal
(Brachycera)
99,942
TOTAL 152,244
209
210 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
REFERENCES
Abul-Hab, J. K. and Y. G. Salman. 1999. Human urino-
genital myasis by Psychoda.Saudi Medical Journal 20:
635–636.
Adler, P. H., D. C. Currie, and D. M. Wood. 2004. The Black
Flies (Simuliidae) of North America. Cornell University Press,
Ithaca, New York. xv +941 pp.
Akam, M. 1998. Ho x genes, homeosis and the evolution of seg-
ment identity: no need for hopeless monsters. International
Journal of Developmental Biology 42: 445–451.
Ali, A. 1991. Perspectives on management of pestiferous Chi-
ronomidae, an emerging global problem. Journal of the
American Mosquito Control Association 7: 260–281.
Aluja, M. and A. L. Norrbom. 1999. Preface. Pp. i –iii In
M. Aluja and A. L. Norrbom (eds). Fruit Flies (Tephritidae):
Phylogeny and Evolution of Behavior. CRC Press, Boca Raton,
Florida. xvi +944 pp.
Alverson, A. J., G. W. Courtney, and M. R. Luttenton. 2001.
Niche overlap of sympatric Blepharicera (Diptera: Blephar-
iceridae) larvae from the southern Appalachian Mountains.
Journal of the North American Benthological Society 20:
564–581.
Andersson, H. 1970. Notes on north European Lonchoptera
(Dipt., Lonchopteridae) with lectotype designations. Ento-
mologisk Tidskrift 91: 42–45.
Armitage, P. D., P. S. Cranston, and L. C. V. Pinder (eds).
1995. The Chironomidae: Biology and Ecology of Non-Biting
Midges. Chapman and Hall, London. xii +572 pp.
Ashburner, M. and C. M. Bergman. 2005. Drosophila mela-
nogaster: a case study of a model genomic sequence and
its consequences. Genome Research 15: 1661–1667.
Axtell, R. C. and J. J. Arends. 1990. Ecol ogya ndm anagement
of arthropod pests of poultry. Annual Review of Entomology
35: 101–126.
Baker, C. H. 2002. Dipteran glow-worms: Marvellous mag-
gots weave magic for tourists. Biodiversity 3 (4): 23.
Barbour, M. T., J. Gerritsen, B. D. Snyder, and J. B. Stribling.
1999. Rapid Bioassessment Protocols for Use in Streams
and Wadable Rivers: Periphyton, Benthic Macroinvertebrates
and Fish. United States Environmental Protection Agency,
Washington, DC. 202 pp.
Bar-Cohen, Y. 2005. Biomimetics: mimicking and inspired-
by biology. Proceedings of the EAPAD Conference,
SPIE Smart Structures and Materials Symposium, Paper
#5759-02. http://ndeaa.jpl.nasa.gov/nasa-nde/lommas/
papers/SPIE-05-Biomimetics.pdf [Accessed 15 November
2007].
Barnes, E. 2005. Diseases and Human Evolution. University of
New Mexico Press, Albuquerque. xii +484 pp.
Barraclough, D. 2006. Bushels of bots: Africa’s largest fly is
getting a reprieve from extinction. Natural History [2006],
June: 18–21.
Baumgartner, D. L. 1993. Review of Chrysomya rufifacies
(Diptera: Calliphoridae). Journal of Medical Entomology 30:
338–352.
Beaman, R. S., P. J. Decker, and J. H. Beaman. 1988. Pollina-
tion of Rafflesia (Rafflesiaceae). American Journal of Botany
75: 1148–1162.
Bennett, F. D. 1969. Tachinid flies as biological control agents
for sugar cane moth borers. Pp. 117–148. In J. R. Williams,
J. R. Metcalfe, R. W. Mungomery,and R. Mathes (eds). Pests
of Sugar Cane. Elsevier, Amsterdam. xiii +568 pp.
Berg, C. O. and Knutson, L. 1978. Biology and systemat-
ics of the Sciomyzidae. Annual Review of Entomology 23:
239–258.
Bess, H. A. and F. H. Haramoto. 1972. Biological control of
pamakani Eupatorium adenophorum, in Hawaii by a tephritid
gall fly, Procecidochares utilis. 3. Status of the weed, fly and
parasites of the fly in 1961–71 versus 1950–57. Pro-
ceedings of the Hawaiian Entomological Society 21: 165–178.
Binot, M., R. Bless, P. Boye, H. Gr ¨uttke, and P. Prescher (eds).
1998. Rote Liste gef ¨ahrdeter Tiere Deutschlands. Schriften-
reie f¨
ur Landschaftspflege und Natursch ¨
utz 55: 1–434.
Bisley, G. G. 1972. A case of intraocular myiasis in man due
to the first stage larva of the oestrid fly Gedoelstia spp. East
African Medical Journal 49: 768–771.
Black, R. W. and M. D. Munn. 2004. Using macroinverte-
brates to identify biota–land cover optima at multiple scales
in the Pacific Northwest, USA. Journal of the North American
Benthological Society 23: 340–362.
Blake, W. 1794. Songs of Experience. London, 17 pls. [Pub-
lishedbytheauthor.HerefromD.V.Erdman(ed).
1988. Electronic edition of the complete poetry and prose
of William Blake. Newly revised edition. http://virtual
.park.uga.edu/wblake/eE.html [Accessed 1 March 2008].
Boettner, G. H., J. S. Elkinton, and C. J. Boettner. 2000. Effects
of a biological control introduction on three nontarget
native species of saturniid moths. Conservation Biology 14:
1798–1806.
Bohart, G. E. 1972. Management of wild bees for pollination
of crops. Annual Review of Entomology 17: 287– 312.
Bohart, G. E., W. P. Stephen, and R. K. Eppley. 1960. The
biology of Heterostylum robustum, a parasite of the alkali
bee. Annals of the Entomological Society of America 53:
425–435.
Bonada, N., N. Prat, V. H. Resh, and B. Statzner. 2006. Devel-
opments in aquatic insect biomonitoring: a comparative
analysis of different approaches. Annual Review of Entomol-
ogy 51: 495–523.
Bondari, K. andD. C. Sheppard. 1981. Soldier fly larvae as feed
in commercial fish production. Aquaculture 24: 103–109.
Bonduriansky, R. and R. J. Brooks. 1998. Male antler flies
(Protopiophila litigata; Diptera: Piophilidae) are more
selective than females in mate choice. Canadian Journal
of Zoology 76: 1277–1285.
Borkent, A. 2004. Ceratopogonidae. Pp. 113–126. In
W. C. Marquardt (ed). Biology of Disease Vectors. Second
edition. Elsevier Academic Press, San Diego. xxiii +785 pp.
Borkent, C. J. and L. D. Harder. 2007. Flies (Diptera) as polli-
nators of two dioecious plants: behaviour and implications
for plant mating. Canadian Entomologist 139: 235–246.
Biodiversity of Diptera 211
Brault, V., M. C. Reedy, U. Sauder, R. A. Kammerer, U. Aebi,
and C. Schoene nberger. 1999. Substitution of flight muscle-
specific actin by human (beta)-cytoplasmic actin in the
indirect flight muscle of Drosophila.Journal of Cell Science
112: 3627–3639.
Brogiolo, W., H. Stocker, T. Ikeya, F. Rintelen, R. Fernandez,
and E. Hafen. 2001. An evolutionarily conserved function
of the Drosophila insulin receptor and insulin-like peptides
in growth control. Current Biology 11: 213–221.
Brown, B. V. 2001. Flies, gnats, and mosquitoes. Pp.
815–826. In S. A. Levin (ed). Encylopedia of Biodiversity.
Academic Press, London.
Brown, B. V. 2004. Revision of the subgenus Udamochiras of
Melaloncha bee-killing flies (Diptera: Phoridae: Metopini-
nae). Zoological Journal of the Linnean Society 140: 1–42.
Brown, B. V. 2008. Phorid Net. http://www.phorid.net/
[Accessed 25 February 2008].
Bruce-Chwatt, L. J. 1988. History of malaria from prehisto-
ry to eradication. Pp. 1–59. In W. H. Wernsdorfer and
I. McGregor (eds). Malaria. Principles and Practice of Malar-
iology. Volume 1. Churchill Livingstone, Edinburg h. 912 pp.
Burger, J. F., J. R. Anderson, and M. F. Knudsen. 1980. The
habits and life history of Oedoparena glauca (Diptera: Dry-
omyzidae), a predator of barnacles. Proceedings of the
Entomological Society of Washington 82: 360–377.
Callaway, E. 2007. Dengue fever climbs the social ladder.
Nature 448: 734–735.
Campbell, H. D., S. Fountain, I. G. Young, C. Claudianos,
J. D. Hoheisel, K. S. Chen, and J. R. Lupski 1997. Genomic
structure, evolution, and expression of human FLII, a gel-
solin and leucine-rich-repeat family member: overlap with
LLGL. Genomics 42: 46–54.
Carey, J. R., N. Papadopoulos, N. Kouloussis,B. Katsoyannos,
H.-G. M ¨uller, J.-L. Wang, and Y.-K. Tseng. 2006. Age-
specific and lifetime behavior patterns in Drosophila
melanogaster and the Mediterranean fruit fly, Ceratitis capi-
tata.Experimental Gerontology 41: 93–97.
Carroll, S. B. 1995. Homeotic genes and evolution of arthro-
pods and chordates. Nature 376: 479–485.
Castle, W. E. 1906. Inbreeding, cross-breeding and sterility in
Drosophila.Science 23: 153.
Chan, W. P. and M. H. Dickinson 1996. In vivo length oscil-
lations of indirect flight muscles in the fruit fly Drosophila
virilis.Journal of Experimental Biology 199: 2767–2774.
Clement, S. L., B. C. Hellier, L. R. Elberson, R. T. Staska, and
M. A. Evans. 2007. Flies (Diptera: Muscidae: Calliphori-
dae) are efficient pollinators of Allium ampeloprasum L.
(Alliaceae) in field cages. Journal of Economic Entomology
100: 131–135.
Clements, A. N. 1992. The Biology of Mosquitoes.Volume1.
Development, Nutrition, and Reproduction. Chapman and
Hall, London. 509 pp.
Cleworth, M. A., R. M. Coombs, and D. H. Davies 1996. The
control of Psychoda alternata (Psychodidae) in sewage
biological filters by application of the insect growth-
regulator pyriproxyfen. Water Research 30: 654–662.
Cock, M. J. W. (ed). 1985. A review of biological control of
pests in the Commonwealth Caribbean and Bermuda up to
1982. Commonwealth Institute of Biological Control, Technical
Communication 9: 1–218.
Cohen, D. 2002. This won’t hurt a bit. New Scientist 174
(2337): 21. Also available as ‘Painless needle copies mos-
quito’s stinger’ at: http://www.newscientist.com/news/
news.jsp?id=ns99992121 and http://www.newscientist
.com/article.ns?id=dn2121 [Accessed 14 February 2008].
Colbo, M. H. 1996. Chironomidae from marine coastal envi-
ronments near St Johns, Newfoundland, Canada. Hydrobi-
ologia 318: 117–122.
Colless, D. H. and D. K. McAlpine. 1991. Diptera (flies). Pp.
717–786. In Commonwealth Scientific and Industrial Research
Organisation (Division of Entomology). The Insects of Australia.
Second Edition. Volume 2. Cornell University Press, Ithaca,
NY. vi +543–1137 pp.
Conley, K. E. and S. L. Lindstedt. 2002. Energy-saving mech-
anisms in muscle: the minimization strategy. Journal of
Experimental Biology 205: 2175–2181.
Cook, D. F., I. R. Dadour, and N. J. Keals. 1999. Stable fly,
house fly (Diptera: Muscidae), and other nuisance fly
development in poultry litter associated with horticultural
crop production. Journal of Economic Entomology 92:
1352–1357.
Coupland, J. and G. Baker. 1995. The potential of several
species of terrestrial Sciomyzidae as biological control
agents of pest helicid snails in Australia. Crop Protection
14: 573–576.
Courtney, G. W. 1990. Cuticular morphology of larval moun-
tain midges (Diptera: Deuterophlebiidae): implications for
the phylogenetic relationships of Nematocera. Canadian
Journal of Zoology 68: 556–578.
Courtney, G. W. 1991. Life history patterns of Nearctic moun-
tain midges (Diptera: Deuterophlebiidae).Journal ofthe North
American Benthological Society 10: 177–197.
Courtney, G. W. 1994. Biosystematics of the Nymphomyiidae
(Insecta: Diptera): life history, morphology and phyloge-
netic relationships. Smithsonian Contributions to Zoology
550: 41.
Courtney, G. W. 2000a. Revision of the net-winged midges of
the genus Blepharicera Macquart (Diptera: Blephariceridae)
of eastern North America. Memoirs of the Entomological
Society of Washington 23: 1–99.
Courtney, G. W. 2000b. A.1. Family Blephariceridae. Pp.
7–30. In L. Papp and B. Darvas (eds). Contributions to a Man-
ual of Palaearctic Diptera. Appendix. Science Herald,Budapest.
604 pp.
Courtney, G. W. and R. W. Merritt. 2008. Aquatic Diptera:
Part one: larvae of aquatic Diptera. Pp. 687–722. In
R. W. Merritt, K. W. Cummins, and M. B. Berg (eds). An
Introduction to the Aquatic Insects of North America.
Fourth Edition. Kendall/Hunt Publishing, Dubuque, Iowa.
1158 pp.
Courtney, G. W., R. W. Merritt, K. W. Cummins, M. B. Berg,
D. W. Webb, and B. A. Foote. 2008. Ecological and
212 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
distributional data for larval aquatic Diptera. Pp. 747–771.
In R. W. Merritt, K. W. Cummins, and M. B. Berg (eds).
An Introduction to the Aquatic Insects of North America.
Fourth Edition. Kendall/Hunt Publishing, Dubuque, Iowa.
1158 pp.
Courtney, G. W., B. J. Sinclair, and R. Meier. 2000. 1.4. Mor-
phology and terminology of Diptera larvae. Pp. 85–161.
In L. Papp and B. Darvas (eds). Contributions to a Manual of
Palaearctic Diptera. Volume I, General and Applied Dipterology.
Science Herald, Budapest. 978 pp.
Craig, D. A. and D. C. Currie. 1999. Phylogeny of the central-
western Pacific subgenus Inseliellum (Diptera: Simuliidae).
Canadian Journal of Zoology 77: 610–623.
Cranston, P. S. and S. Dimitriadis. 2005. Semiocladius: taxon-
omy and ecology of an estuarine midge (Diptera, Chirono-
midae, Orthocladiinae). Australian Journal of Entomology 44:
252–256.
Crosby, M. C. 2006. The American Plague: The Untold Story of
Yellow Fever, the Epidemic that Shaped Our History. Berkley
Books, New York. viii +308 pp.
Crosskey, R. W. 1990. The Natural History of Blackflies.John
Wiley and Sons, Chichester. ix +711 pp.
Cumming, J. M. 1994. Sexual selection and the evolution of
dance fly mating systems (Diptera: Empididae; Empidinae).
Canadian Entomologist 126: 907–920.
Cumming, J. M., B. J. Sinclair, andD. M. Wood. 1995. Homol-
ogy and phylogenetic implications of male genitalia
in Diptera – Eremoneura. Entomologica Scandinavica 26:
121–151.
Dahl, C. 1960. Musidoridae (Lonchopteridae) Diptera, Cyclo r-
rhapha from the Azores. Boletim do Museu Municipal do
Funchal 13: 96.
De Buck, N. 1990. Bloembezoek en bestuivingsecologie van
Zweefvliegen (Diptera, Syrphidae) in het bijzonder voor
Belgie. Studiendocumenten Royal Belgian Institute of Natural
Sciences 60: 1–167.
DeBach, P. and D. Rosen. 1991. Biological Control by Natu-
ral Enemies. Second Edition. Cambridge University Press,
Cambridge. 440 pp.
Delir, S., F. Handjani, M. Emad, and S. Ardehali. 1999. Vulvar
myiasis due to Wohlfahrtia magnifica.Clinical and Experimen-
tal Dermatology 24: 279–280.
Dempewolf, M. 2004. Arthropods of Economic Importance –
Agromyzidae of the World. CD-ROM series, ETI/ZMA,
Amsterdam. Also available at http://nlbif.eti.uva.nl/bis/
agromyzidae.php
Desportes, C. 1942. Forcipomyia velow Winn. et Sycorax silacea
Curtis, vecteurs d’Icosiella neglecta (Diesing), filaire com-
mune de la grenouille verte. Annales de ParasitologieHumaine
et Compar´
ee 19: 53–68.
Diggins, T. P. and K. M. Stewart. 1993. Deformities of aquatic
larval midges (Chironomidae: Diptera) in the sediments of
the Buffalo River, New York. Journal of Great Lakes Research
19: 648–659.
Dimitriadis, S. and P. S. Cranston 2007. From the moun-
tains to the sea: assemblage structure and dynamics in
Chironomidae (Insecta: Diptera) in the Clyde River estuar-
ine gradient, New South Wales, south-eastern Australia.
Australian Journal of Entomology 46: 188–197.
Disney, R. H. L. 1991. The aquatic Phoridae (Diptera). Ento-
mologica Scandinavica 22: 171–191.
Disney, R. H. L. 1994a. Continuing the debate relating to
the phylogenetic reconstruction of the Phoridae (Diptera).
Giornale Italiano di Entomologia 7: 103–117.
Disney, R. H. L. 1994b. Scuttle Flies: the Phoridae. Chapman
and Hall, London. xii +467 pp.
Disney, R. H. L. 1998. Family Phoridae. Pp. 51 –79. In L. Papp
and B. Darvas (eds). Contributions to a Manual of Palaearc-
tic Diptera. Volume 3.Higher Brachycera. Science Herald,
Budapest. 880 pp.
Disney, R. H. L. and D. H. Kistner. 1995. Revision of the
Afrotropical Termitoxeniinae (Diptera: Phoridae). Sociobi-
ology 26: 117–225.
Disney, R. H. L. and H. Kurahashi. 1978. A case of urogen-
ital myiasis caused by a species of Megaselia (Diptera:
Phoridae). Journal of Medical Entomology 14: 717.
Dowell, R. V. and L. K. Wange. 1986. Process analysis and
failure avoidance in fruit fly programs. Pp. 43–65. In
M. Mangel, J. R. Carey, and R. E. Plant (eds). Pest Con-
trol: Operations and Systems Analysis in Fruit Fly Man-
agement. NATO Advanced Science Institutes Series G.
Ecological Sciences Springer-Verlag, New York 11 xii +
465 pp.
Downes, W. L., Jr. 1986. The Nearctic Melanomya and rel-
atives (Diptera: Calliphoridae), a problem in calyptrate
classification. Bulletin of the New York State Museum 460: v
+1–35.
Dudley, T. L. and N. H. Anderson. 1987. The biology and
life cycles of Lipsothrix spp. (Diptera: Tipulidae) inhabiting
wood in western Oregon streams. Freshwater Biology 17:
437–451.
Dunn, L. H. 1930. Rearing the larvae of Dermatobia hominis
Linn. in man. Psyche 37: 327–342.
Dwortzan, M. 1997. It’s a Fly! It’s a Bug! It’s a Microplane!
http://www2.technologyreview.com/articles/97/10/
reporter21097.asp [Accessed 1 April 2005].
Eibl, J. and R. Copeland. 2005. [Explanation of a phan-
tom phenomenon.] http://www.sel.barc.usda.gov/Diptera/
youngent/clouds explanation.htm [Accessed 8 February
2008.]
Eigen, M., W. J. Kloft, and G. Brandner. 2002. Transferability
of HIV by arthropods supports the hypothesis about trans-
mission of the virus from apes to man. Naturwissenschaften
89: 185–186.
Elberling, H. and J. M. Olesen. 1999. The structure of a high
latitude plant-flower visitor system: the dominance of flies.
Ecography 22: 314–323.
Ellington, C. P. 1999. The novel a erodynamicsof insect flight:
applications to micro-air vehicles. Journal of Experimental
Biology 202: 3439–3448.
Emerson, P. M., R. L. Bailey, and O. S. Mahdi. 2000. Trans-
mission ecology of the fly Musca sorbens, a putative vector
Biodiversity of Diptera 213
of trachoma. Transactions of the Royal Society of Tropical
Medicine and Hygiene 94: 28–32.
Endress, P. K. 2001. The flowers in extant basal angiosperms
and inferences on ancestral flowers. International Journal of
Plant Sciences 162: 1111–1140.
Englund, R. A., M. G. Wright, and D. A. Polhemus. 2007.
Aquatic insect taxa as indicators of aquatic species rich-
ness, habitat disturbance, and invasive species impacts in
Hawaiian streams. In N. L. Evenhuis and J. M. Fitzsimons
(eds). Biology of Hawaiian streams and estuaries. Bishop
Museum Bulletin in Cultural and Environmental Studies 3:
207–232.
Eschle, J. L. and G. R. DeFoliart. 1965. Control of mink myiasis
caused by the larvae of Wohlfahrtia vigil. Journal of Economic
Entomology 58: 529–531.
Esser, J. R. 1991. Biology of Chrysomya megacephala (Diptera:
Calliphoridae) and reduction of losses to the salted-dried
fish industry in south-east Asia. Bulletin of Entomological
Research 81: 33–41.
Evenhuis, N. L., T. Pape, A. C. Pont, and F. C. Thompson
(eds). 2007. BioSystematics Database of World Diptera,
Version 9.5. http://www.diptera.org/biosys.htm [Accessed
20 January 2008].
Falk, P. 1991. A review of the scarce and threatened flies
of Great Britain (part I). Research and Survey in Nature
Conservation 39: 1–194.
Fast, E. and T. A. Wheeler. 2004. Faunal inventory of
Brachycera (Diptera) in an old growth forest at Mont Saint-
Hilarie, Quebec. Fabreries 29: 1–15.
Farkas, R. and G. Kepes. 2001. Traumatic myiasis of horses
caused by Wohlfahrtia magnifica.Acta Veterinaria Hungarica
49: 311–318.
Farkas, R., Z. Szanto, and M. R. J. Hall. 2001. Traumatic myi-
asis of geese in Hungary. Veterinary Parasitology 95: 45–52.
Ferrar, P. 1976. Macrolarviparous reproduction in Ameni-
inae (Diptera: Calliphoridae). Systematic Entomology 1:
107–116.
Ferrar, P. 1987. A guide to the breeding habits and immature
stages of Diptera Cyclorrhapha. Entomonograph 8: 1–907.
[2 volumes.]
Fessl, B., B. J. Sinclair, and S. Kleindorfer. 2006. The life-cycle
of Philornis downsi (Diptera: Muscidae) parasitizing Darwin’s
finches and its impacts on nestling survival. Parasitology
133: 739–747.
Foote, B. A. (coordinator). 1987. Order Diptera. Pp.
690–915. In F. W. Stehr (ed). Immature Insects. Volume 2.
Kendall/Hunt Publishing, Dubuque, Iowa. 975 pp.
orster, M., S. Klimpel, H. Mehlhorn, K. Sievert, S. Messler,
and K. Pfeffer. 2007. Pilot study on synanthropic flies (e.g.
Musca, Sarcophaga, Calliphora, Fannia, Lucilia, Stomoxys)
as vectors of pathogenic microorganisms. Parasitological
Research 101: 243–246.
Fortini, M. E., M. P. Skupski, M. S. Boguski, and I. K. Hariha-
ran. 2000. A survey of human disease gene counterparts
in the Drosophila genome. Journal of Cell Biology 150:
F23–F29.
Fredeen, F. J. H. and M. E. Taylor. 1964. Borborids (Diptera:
Sphaeroceridae) infesting sewage disposal tanks, with notes
on the life cycle, behaviour and control of Leptocera caenosa
(Rond.). Canadian Entomologist 96: 801–808.
Free, J. B. 1993. Insect Pollination of Crops. Second Edition.
Academic Press, London. xii +684 pp.
Gaimari, S. D. and W. J. Turner. 1996. Methods for rearing
aphidophagous Leucopis spp. (Diptera: Chamaemyiidae).
Journal of Kansas Entomological Society 69: 363–369.
Gagn´e, R. J. 1994. The Gall Midges of the Neotropical Region.
Cornell University Press, Ithaca, New York. xv +352 pp.
Gagn´e, R. J. 1995. Revision of tetranychid (Acarina) mite
predators of the genus Feltiella (Diptera: Cecidomyiidae).
Annals of the Entomological Society of America 88: 16–30.
Gagn´e, R. J., H. Blanco-Metzler, and J. Etienne. 2000. A new
Neotropical species of Clinodiplosis (Diptera: Cecidomyiidae),
an important new pest of cultivated peppers (Capsicum
spp.: Solanaceae). Proceedings of the Entomological Society of
Washington 102: 831–837.
Gagn´e, R. J., A. Sosa, and H. Cordo. 2004. A new Neotropical
species of Clinodiplosis (Diptera: Cecidomyiidae) injurious
to alligatorweed, Alternanthera philoxeroides (Amaran-
thaceae). Proceedings of the Entomological Society of Wash-
ington 106: 305–311.
Gattiker, G., K. V. I. S. Kaler, and M. P. Mintchev. 2005. Elec-
tronic mosquito: designing a semi-invasive microsystem for
blood sampling, analysis and drug delivery applications.
Microsystem Technologies 12: 44–51.
Gillespie, D. R., B. Roitberg, E. Basalyga, M. Johnstone,
G. Opit, J. Rodgers, and N. Sawyer. 1998. Biology and appli-
cation of Feltiella acarisuga (Vallot) (Diptera: Cecidomyiidae)
for biological control of twospotted spider mites on green-
house vegetable crops. Pacific Agri-Food Research Centre
(Agassiz). Technical Report 145. Agriculture and Agri-Food
Canada.
Gold, B. L., K. P. Mathews, and H. A. Burge. 1985. Occupa-
tional asthma caused by sewer flies. American Review of
Respiratory Diseases 131: 949–952.
Goldblatt, P. and J. C. Manning. 2000. The long-proboscid fly
pollination system in southern Africa. Annals of the Missouri
Botanical Garden 87: 146–170.
Golding, W. 1954. Lord of the Flies. Faber and Faber Ltd.,
London. 248 pp.
Gonzalez, L. and B. V. Brown. 2004. New species and records
of Melaloncha (Udamochiras) bee-killing flies (Diptera:
Phoridae). Zootaxa 730: 1–14.
Good, J. A. and M. C. D. Speight. 1996. Saproxylic Invertebrates
and Their Conservation Throughout Europe. Convention on
the Conservation of European Wildlife and Natural Habitats.
Council of Europe, Strasbourg. 58 pp.
Graham, L. C., S. D. Porter, R. M. Pereira, H. D. Dorough, and
A. T. Kelley. 2003. Field releases of the decapitating fly
Pseudacteon curvatus (Diptera: Phoridae) for control of
imported fire ants (Hymenoptera: Formicidae) in Alabama,
Florida, and Tennessee. Florida Entomologist 86: 334–
339.
214 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
Grantham-Hill, C. 1933. Preliminary note on the treatment
of infected wounds with the larva of Wohlfartia [sic]nuba.
Transactions of the Royal Society of Tropical Medicine and
Hygiene 27: 93–98.
Greathead, D. J. 1983. Natural enemies of Nilaparvata lugens
and other leaf- and planthoppers in the tropical agroe-
cosystems and their impact on pest population(s). Pp.
371–383. In J. W. Knight, N. C. Pant, T. S. Robertson, and
M. R. Wilson (eds). First International Workshop on Leafhop-
pers and Planthoppers of Economic Importance. Common-
wealth Institute of Entomology, London. 500 pp.
Greathead, D. J. 1995. The Leucopis spp. (Diptera: Chamae-
myiidae) introduced for biological control of Pineus sp.
(Homoptera: Adelgidae) in Hawaii: implications for control
of Pineus boerneri in Africa. The Entomologist 114: 83–90.
Greathead, D. J. and N. L. Evenhuis. 1997. Family Bombyli-
idae. Pp. 487– 512. In L. Papp and B. Darvas (eds). Contribu-
tions to a Manual of Palaearctic Diptera. Volume 2. Nematocera
and Lower Brachycera. Science Herald. Budapest. 592 pp.
Greenberg, B. 1971. Flies and Diseases. Volume 1. Ecology
Classification and Biotic Associations. Princeton University
Press, Princeton. 856 pp.
Greenberg, B. 1973. Flies and Disease. Volume II. Biology and
Disease Transmission. Princeton University Press, Prince-
ton. 447 pp.
Griffiths, G. C. D. 1972.The Phylogenetic Classification of D´
ıptera
Cyclorrhapha with Special Reference to the Structure of the
Male Postabdomen. Series Entomologica 8. Dr. W. Junk, The
Hague. 340 pp.
Grimaldi, D. and J. M. Cumming. 1999. Brachyceran Diptera
in Cretaceous ambers and Mesozoic diversification of the
Eremoneura. Bulletin of the American Museum of Natural
History 239: 1–124.
Grimaldi, D. and M. S. Engel. 2005. Evolution of the Insects.
Cambridge University Press, Cambridge. 755 pp.
Grisi, L., C. L. Massard, G. E. Moya-Borja, and J. B. Pereira.
2002. Impacto econ ˆomico das principais ectoparasitoses
em bovinos no Brasil. Hora Veterin´
aria 21: 8–10.
Grunin, K. J. 1973. The first discovery of larvae of the mam-
moth bot-fly Cobboldia (Mamontia, subgen. n.) russanovi sp.n.
(Diptera, Gasterophilidae). ´
Entomologicheskoe Obozrenie 52:
228–233. [In Russian with English subtitle; English trans-
lation: Entomological Rev iew, Washington 52: 165–169].
Guimar˜aes, J. H. 1977. A systematic revision of the Mesem-
brinellidae, stat. nov. (Diptera, Cyclorrhapha). Arquivos de
Zoologia, S˜
ao Paulo 29: 1–109.
Guimar˜aes, J. H. and N. Papavero. 1999. Myiasis in Man and
Animals in the Neotropical Region; Bibliographic Database.
Flˆeiade/FAPESP, S˜ao Paulo, Brazil. 308 pp.
Guimar˜aes, J. H., N. Papavero, and A. Prado. 1983. As mi´
ıases
na regi˜ao Neotropical (identificac¸˜ao, biologia, bibliografia).
Revista Brasileira de Zoologia 1: 239–416.
Haines, C. P. and D. P. Rees. 1989. A field guide to the
types of insects and mites infesting cured fish. FAO
Fisheries Technical Paper 303: 1–33. [Also available at
http://www.fao.org/documents/ [Accessed 3 June 2005].
Hall, M. J. R. 1997. Traumatic myiasis of sheep in Europe: a
review. Parasitologia 39: 409–413.
Hall, M. J. R. and R. Wall. 1995. Myiasis of humans and
domestic animals. Advances in Parasitology 35: 257–334.
Hansen, M. 1983. Yuca (Yuca, Cassava). Pp. 114–117. In
D. Janzen (ed). Costa Rican Natural History. University of
Chicago Press, Chicago. xi +816 pp.
Harbach, R. E. 2007. The Culicidae (Diptera): a review of
taxonomy, classification and phylogeny. Zootaxa 1668:
591–638.
Harris, P. 1989. The use of Tephritidae for the biological
control of weeds. Biocontrol News and Information 10: 7–16.
Harrison, G. 1978. Mosquitoes, Malaria and Man: a History
of the Hostilities Since 1880. First Edition. John Murray,
London. viii +314 pp.
Hashimoto, H. 1976. Non-biting midges of marine habitats
(Diptera: Chironomidae). Pp. 377 –414. In L. Cheng (ed).
Marine Insects. North-Holland Publishing Co., Amsterdam,
The Netherlands. 581 pp.
Haslett, J. R. 1988. Qualit¨atsbeurteilung alpiner Habitate:
Schwebfliegen (Diptera: Syrphidae) als Bioindikatoren ur
Auswirkungen des intensiven Skibetriebes auf alpinen
Wiesen in ¨
Osterreich. Zoologischer Anzeiger 220: 179–
184.
Hawaii Biological Survey. 2002. Hawaii’s Extinct Species
Insects. http://hbs.bishopmuseum.org/endangered/
ext-insects.html [Accessed 4 March 2008].
Heath, A. C. G. 1982. Beneficial aspects of blowflies (Diptera:
Calliphoridae). New Zealand Entomologist 7: 343–348.
Hennig, W. 1971. Neue untersuchungen ueber die fam-
ilien der Diptera Schizophora (Diptera: Cyclorrhapha).
Stuttgarter Beitraege zur Naturkunde, Serie A (Biologie)
226: 1–76.
Hennig, W. 1973. Diptera (Zweifl ¨ugler). Handbuch der Zoologie
(Berlin) 4: 1–200.
Hern´andez, F. O. and A. A. Guti´errez. 2001. Avoiding Pseu-
dohypocera attacks (Diptera: Phoridae) during the artificial
propagation of Melipona beecheii colonies (Hymenoptera:
Apidae: Meliponini). Folia Entomologica Mexicana 40:
373–379.
Hide, G. 1999. History of sleeping sickness in East Africa.
Clinical Microbiology Reviews 12: 112–125.
Higley, L. G. and N. H. Haskell. 2001. Insect development
and forensic entomology. Pp. 287–302. In J. H. Byrd and
J. L. Castner (eds). Forensic Entomology. CRC Press, Boca
Raton, Florida. 418 pp.
Hill, D. S. 1963. The life history of the British species of
Ornithomya (Diptera: Hippoboscidae). Transactions of the
Royal Entomological Society of London 115: 391–407.
Hoffmann, M. P., and A. C. Frodsham. 1993. Natural Enemies
of Vegetable Insect Pests. Cooperative Extension, Cornell
University, Ithaca, New York. 63 pp.
Hogue, C. L. 1981. Blephariceridae. Pp. 191–197. In
J. F. McAlpine, B. V. Petersen, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). Manual of
Nearctic Diptera, Volume 1. Research Branch, Agriculture
Biodiversity of Diptera 215
Canada Monograph 27. Supply & Services Canada, Hull,
Quebec.
Holloway, B. A. 1976. Pollen-feeding in hover-flies (Diptera:
Syrphidae). New Zealand Journal of Zoology 3: 339–350.
Holston, K. C. 2005. Evidence for community structure and
habitat partitioning by stiletto flies (Diptera: Therevidae) at
the Guadalupe-Nipomo Dune System, California. Journal of
Insect Science 5 (42): 1–17.
Horgan, F. G., J. H. Myers, and R. Van Meel. 1999. Cyzenis
albicans (Diptera: Tachinidae) does not prevent the out-
break of winter moth (Lepidoptera: Geometridae) in birch
stands and blueberry plots on the lower mainland of British
Columbia. Environmental Entomology 28: 96–107.
Horiuchi, J. and M. Saitoe. 2005. Can flies shed light on
our own age-related memory impairment. Ageing Research
Reviews 4: 83–101.
ovemeyer, K. 2000. 1.11. Ecology of Diptera. Pp. 437–489.
In L. Papp and B. Darvas (eds). Contributions to a Manual of
Palaearctic Diptera. Volume 1. General and Applied Dipterology.
Science Herald, Budapest. 978 pp.
Howard, J. 2001. Nuisance flies around a landfill: patterns
of abundance and distribution. Waste Management and
Research 19: 308–313.
Hull, F. M. 1962. Robber flies of the World. The genera of the
family Asilidae. Bulletin of the United States National Museum
224: 1–907.
Hussey, N. W. 1960. Biology of mushroom phorids. Mushroom
Science 4: 260–269 [1959].
Iori, A., B. Zechini, L. Cordier, E. Luongo, G. Pontuale, and
S. Persichino. 1999. A case of myiasis in man due to
Wohlfahrtia magnifica (Schiner) recorded near Rome. Para-
sitologia 41: 583–585.
IUCN. 2007. 2007 IUCN Red List of Threatened Species.
http://www.iucnredlist.org/ [Accessed 4 March 2008].
Jackman, R., S. Nowicki, D. J. Aneshansley, and T. Eisner.
1983. Predatory capture of toads by fly larvae. Science
222: 515–516.
Jackson, D. S. and B. G. Lee. 1985. Medfly in California
1980–1982. Bulletin of the Entomological Society of America
1985: 29–37.
James, M. T. 1935. The genus Hermetia in the United States
(Diptera: Stratiomyidae). Bulletin of Brooklyn Entomological
Society 30: 165–170.
Jehl, J. R., Jr. 1986. Biology of red-necked phalaropes (Phalaro-
pus lobatus) at the western edge of the Great Basin in fall.
Great Basin Naturalist 46: 185–197.
Juliano, S. A. and L. P. Lounibos. 2005. Ecology of invasive
mosquitoes: effects on resident species and on human
health. Ecology Letters 8: 558–574.
Kaneshiro, K. Y. 1997. R. C. L. Perkins’ legacy to evolution-
ary research on Hawaiian Drosophilidae (Diptera). Pacific
Science 51: 450–461.
Karouna-Renier, N. K. and J. P. Zehr. 1999. Ecological impli-
cations of molecular biomarkers: assaying sub-lethal stress
in the midge Chironomus tentansusing heat shock protein 70
(HSP-70) expression. Hydrobiologia 401: 255–264.
Kearns, C. A. 1992. Anthophilous fly distribution across
an elevation gradient. American Midland Naturalist 127:
172–182.
Kearns, C. A. 2001. North American dipteran pollinators:
assessing their value and conservation status. Conservation
Ecology 5 (1): 5. http://www.consecol.org/vol5/iss1/art5/
[Accessed 15 November 2007].
Kearns, C. A. 2002. Flies and flowers: an enduring partner-
ship. Wings (The Xerces Society) 25 (2): 3–8.
Keilin, D. 1919. On the life history and larval anatomy of
Melinda cognata Meigen, parasitic in the snail Helicella
(Heliomanes) virgata da Costa, with an account of the other
Diptera living upon molluscs. Parasitology 11: 430–454.
Kevan, P. G. 1972. Insect pollination of high arctic flowers.
Journal of Ecology 60: 831–847.
Kevan, P. G. 2002. Flowers, pollination, and the associated
diversity of flies. Biodiversity 3 (4): 16–18.
Kiesenwetter, E. A. 1857. Chlorops nasuta Meig. in grossen
Schw¨armen beobachtet. Berliner Entomologische Zeitschrift
1: 171.
Klemm, D. J., K. A. Blocksom, F. A. Fulk, A. T. Herlihy,
R. M. Hughes, P. R. Kaufman, D. V. Peck, J. L. Stoddard,
W. T. Theony, M. B. Griffith, and W. S. Davis 2003. Devel-
opment and evaluation of a Macroinvertebrate Biotic
Integrity Index (MBII) for regionally assessing Mid-
Atlantic Highlands streams. Environmental Management 31:
656–669.
Koenig, D. P. and C. W. Young. 2007. First observation of
parasitic relations between big-headed flies of the genus
Nephrocerus (Diptera: Pipunculidae) and crane flies of the
genus Tipula (Diptera: Tupulidae: Tipulinae), with larvaland
puparial descriptions of Nephrocerus atrapilus Skevington.
Proceedings of the Entomological Society of Washington 109:
52–65.
Kotrba, M. 2004. Fliegenschw ¨arme im s ¨udbayerischen
Seengebiet (Diptera: Chloropidae). Nachrichtenblatt der Bay-
erischen Entomologen 53: 32.
KPLC, Lakes Charles, Louisiana. 2004. West Nile virus
cost LA $20 million. http://www.kplctv.com/global/
story.asp?s=2358859andClientType=Printable [Accessed
20 February 2008].
Krafsur, E. S. and R. D. Moon. 1997. Bionomics of the face
fly, Musca autumnalis.Annual Review of Entomology 42:
503–523.
uhne, S. 2000. R ¨auberische Fliegen der Gattung Coenosia
Meigen, 1826 (Diptera: Muscidae) und die M ¨oglichkeit
ihres Einsatzes bei der biologischen Sch¨adlingsbek ¨ampfung.
Studia Dipterologica, Supplement 9: 1–78.
Kutty, S., M. V. Bernasconi, F. Sifner, and R. Meier. 2007.
Sensitivity analysis, molecular systematics, and naturalhis-
tory evolution of Scathophagidae (Diptera: Cyclorrhapha:
Calyptratae). Cladistics 23: 64–83.
Labandeira, C. C. 1998. How old is the flower and the fly.
Science 280: 85–88.
Lagac´e-Wiens, P. R. S., R. Dookeran, S. Skinner, R. Leicht,
D. D. Colwell, and T. D. Galloway. 2008. Human
216 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
ophthalmomyiasis interna caused by Hypoderma tarandi,
Northern Canada. Emerging Infectious Diseases 14: 64–66.
Langelaan, G. 1957. The Fly. Playboy Magazine 42: 16– 18.
Larson, B. M. H., P. G. Kevan, and D. W. Inouye. 2001. Flies
and flowers: taxonomic diversity of anthophiles and polli-
nators. Canadian Entomologist 133: 439–465.
Laurence, B. R. 1953. Some Diptera bred from cow dung.
Entomologist’s Monthly Magazine 89: 281–283.
Laurence, B. R. 1954. The larval inhabitants of cow pats.
Journal of Animal Ecology 23: 234–260.
Laurence, B. R. 1955. Flies associated with cow dung. Ento-
mologist’s Record and Journal of Variation 67: 123–126.
Leclercq, M. 1990. Utilisation de larves de dipt`eres – maggot
therapy – en edicine: historique et actualit´e. Bulletin
et Annales de la Soci´
et´
e royale Belge d’Entomologique 126:
41–50.
Lenat, D. R. 1993a. Using mentum deformities of Chironomus
larvae to evaluate the effects of toxicity and organic loading
in streams. Journal of the North American Benthological Society
12: 265–269.
Lenat, D. R. 1993b . A biotic index for the southeastern United
States: derivation and list of tolerance values, with criteria
for assigning water-quality ratings. Journal of the North
American Benthological Society 12: 279–290.
Letzner, K. W. 1873. Schw ¨arme der Chlorops ornata Meigen.
Jahresbericht der Schlesischen Gesellschaft f ¨
ur Vaterl¨
andische
Cultur 50: 193–199.
Leevers, S. J. 2001. Growth control: invertebrate insulin sur-
prises! Current Biology 11: 209–212.
Lewis, E. B. 1998. The bithorax complex: the first fifty years.
International Journal of Developmental Biology 42: 403–415.
Lindegaard, C. and P. M. J´onasson. 1979. Abundance, pop-
ulation dynamics and production of zoobenthos in Lake
yvatn, Iceland. Oikos 32: 202 –227.
Lindsay, D. R. and H. I. Scudder. 1956. Nonbiting flies and
disease. Annual Review of Entomology 1: 323–346.
Linevich, A. A. 1971. The Chironomidae of Lake Baikal.
Limnologica (Berlin) 8: 51–52.
Linley, J. R. 1976. Biting midges of mangrove swamps and
salt-marshes (Diptera: Ceratopogonidae). Pp. 335–376. In
L. Cheng (ed). Marine Insects. N orth-Holland Publishing Co.,
Amsterdam, The Netherlands. 581 pp.
Lonsdale, O. and S. Marshall. 2004. Clusiidae. Ver-
sion 10, December 2004. http://tolweb.org/Clusiidae/
10628/2004.12.10 In The Tree of Life Web Project:
http://tolweb.org/ [Accessed on 7 March 2008].
Lyons, M. 1992. The Colonial Disease: a Social History of Sleeping
Sickness in Northern Zaire, 1900– 1940. Cambridge Univer-
sity Press, New York. 335 pp.
Maharaj, R., C. C. Appleton, and R. M. Miller. 1992. Snail pre-
dation by larvae of Sepedon scapularis Adams (Diptera:
Sciomyzidae), a potential biocontrol agent of snail inter-
mediate hosts of schistosomiasis in South Africa. Medical
and Veterinary Entomology 6: 183–187.
Maret, T. R., D. J. Cain, D. E. MacCoy, and T. M. Short. 2003.
Response of Benthic invertebrate assemblages to metal
exposure and bioaccumulation associated with hard-rock
mining in northwestern streams, USA.Journal of the North
American Benthological Society 22: 598– 620.
Marshall, A. G. 1970. The life cycle of Basilia hispida Theodor
1967 (Diptera: Nycteribiidae) in Malaysia. Parasitology 61:
1–18.
Marshall, S. A. and O. W. Richards. 1987. Sphaeroceri-
dae. Pp. 993–1006. In J. F. McAlpine, B. V. Peterson,
G. E. Shewell, H. J. Teskey, J. R. Vockeroth,and D. M. Wood
(coordinators). Manual of Nearctic Diptera.Volume2.
Research Branch, Supply & Services Canada, Hull, Quebec,
Agricultural Canada Monograph 28.
Masoodi, M. and K. Hosseini. 2004. External ophthalmomyi-
asis caused by sheep botfly (Oestrus ovis) larva: a report of 8
cases. Archives of Iranian Medicine 7: 136–139.
Matthiessen, J., L. Hayles, and M. Palmer. 1984. An assess-
ment of some methods for the bioassay of changes in cattle
dung as insect food, using the bush fly, Musca vetustis-
sima, Walker (Diptera: Muscidae). Bulletin of Entomological
Research 74: 463–467.
Maughan, D., J. Moore, J. Vigoreaux, B. Barnes, and
L. A. Mulieri. 1998. Work production and work absorp-
tion in muscle strips from vertebrate cardiac and insect
flight muscle fibers. Advances in Experimental and Medical
Biology 453: 471–480.
McAlpine, D. K. 2007. Review of the Borboroidini or wom-
bat flies (Diptera: Heteromyzidae), with reconsideration of
the status of families Heleomyzidae and Sphaeroceridae,
and descriptions of femoral gland baskets. Records of the
Australian Museum 59: 143–219.
McAlpine, J. F. 1973. A fossil ironomyiid fly from Canadian
amber (Diptera: Ironomyiidae). Canadian Entomologist 105:
105–111.
McAlpine, J. F. 1989. Phylogeny and classification of the
Muscomorpha. Pp. 1397– 1518. In J. F. McAlpine and
D. M. Wood (coordinators). Manual of Nearctic Diptera.Vol-
ume 3. Research Branch, Supply & Services Canada, Hull,
Quebec, Agricultural Canada Monograph 32.
McAlpine, J. F., B. V. Peterson, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). 1981.
Manual of Nearctic Diptera. Volume 1. Research Branch, Sup-
ply & Services Canada, Hull, Quebec, Agriculture Canada
Monograph 27. 674 pp.
McAlpine, J. F., B. V. Peterson, G. E. Shewall, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). 1987.
Manual of Nearctic Diptera.Volume 2. Research Branch, Sup-
ply & Services Canada, Hull, Quebec, Agriculture Canada
Monograph 28. 675– 1332 pp.
McElravy, E. P. and V. H. Resh. 1991. Distribution and sea-
sonal occurrence of the hyporheic fauna in a northern
California stream. Hydrobiologia 220: 233–246.
McLean, I. F. G. 2000. Beneficial Diptera and their role in
decomposition. Pp. 491–517. In L. Papp and B. Darvas
(eds). Contributions to a Manual of Palaearctic Diptera.Vol-
ume 1. General and Applied Dipterology. Science Herald,
Budapest. 978 pp.
Biodiversity of Diptera 217
McPheron, B. A. and G. J. Steck (eds). 1996. Fruit Fly Pests: A
World Assessment of Their Biology and Management.St.Lucie
Press, Delray Beach. 586 pp.
Meier, R., M. Kotrba, and P. Ferrar. 1999. Ovoviviparity and
viviparity in the Diptera. Biological Reviews 74: 199–258.
Mellor, P. S. and J. Boorman. 1995. The transmission and
geographical spread of African horse sickness and blue-
tongue viruses. Annals of Tropical Medicine and Parasitology
89: 1–15.
Menzel, F. and W. Mohrig. 1999. Revision d er pal¨aarktischen
Trauerm ¨ucken (Diptera: Sciaridae). Studia Dipterologica,
Supplement 6: 1–761.
Merritt, R. W., G. W. Courtney, and J. B. Keiper. 2003.
Diptera. Pp. 324–340. In V. H. Resh and R. T. Card ´e(eds).
Encyclopedia of Insects. Academic Press, London.
Michailova, P., N. Petrova,G. Sella, L. Ramella, and S. Bovero.
1998. Structural-functional rearrangements in chromo-
some G in Chironomus riparius (Diptera, Chironomidae)
collected from a heavy metal-polluted area near Turin,
Italy. Environmental Pollution 103: 127–134.
MicrobiologyBytes. 2007. Malaria. http://www
.microbiologybytes.com/introduction/Malaria.html [Last
updated 26 April 2007, accessed 10 February 2008].
Mitra, B. and D. Banerjee 2007. Fly pollinators: assessing their
value in biodiversity conservation and food security in India.
Records of the Zoological Survey India 107 (Part 1): 33–48.
Morita, S. I. 2007. Pangonid.net. http://pangonid.net/pages/
philoimages/pages/image1.html [Accessed 3 March 2008].
Morozova, T. V., R. R. H. Anholt, and T. F. C. Mackay. 2006.
Transcriptional response to alcohol exposure in Drosophila
melanogaster.Genome Biology 7 (10): R95.
Mostovski, M. B. 1995. New taxa of ironomyiid flies (Diptera,
Phoromorpha, Ironomyiidae) from Cretaceous deposits of
Siberia and Mongolia. Paleontologicheski Zhurnal 4: 86–103.
Moulton, J. K. and B. M. Wiegmann. 2007. The phyloge-
netic relationships of flies in the superfamily Empidoidea
(Insecta: Diptera). Molecular Phylogenetics and Evolution 43:
701–713.
Moulton, S. R., II, J. L. Carter, S. A. Grotheer, T. F. Cuffney,
and T. M. Short. 2000. Methods of analysis by the U.S. Geolog-
ical Survey National Water Quality Laboratory – processing,
taxonomy, and quality control of benthic macroinvertebrates.
U.S. Geological Survey Open-File Report 00–212.United
States Geological Survey, Reston, Virginia.
Mounier, N., M. Gouy, D. Mouchiroud, and J. C. Prudhomme.
1992. Insect muscle actins differ distinctly from inverte-
brate and vertebrate cytoplasmic actins. Journal of Molecular
Evolution 34: 406–415.
Mulla, M. S. 1965. Biology and control of Hippelates eye gnats.
Proceedings and Papers of the Annual Conference of the Califor-
nia Mosquito Control Association 33: 26–28.
Nartshuk, E. and S. Pakalniˇskis. 2004. Contribution to the
knowledge of the family Chloropidae (Diptera, Muscomor-
pha) of Lithuania. Acta Zoologica Lituanica 14 (2): 56– 66.
Nartshuk, E. P. 1997. Family Acroceridae. Pp. 469–485. In
L. Papp and B. Darvas (eds). Contributions to a Manual of
Palaearctic Diptera.Volume 2. Nematocera and Lower Brachyc-
era. Science Herald, Budapest. 592 pp.
Nartshuk, E. P. 2000. Periodicity of outbreaks of the predatory
fly Thaumatomyia notata Mg. (Diptera, Chloropidae) and its
possible reasons. Entomologicheskoe Obozrenie 79: 771–781.
[In Russian; English translation in Entomological Review 80:
911–918.]
Nayduch, D., G. P. Noblet, and F. J. Stutzenberger 2002. Vec-
tor potential of houseflies for the bacterium Aeromonas
caviae.Medical and Veterinary Entomology 16: 193–198.
Neff, J. L., B. B. Simpson, N. L. Evenhuis, and G. Dieringer.
2003. Character analysis of adaptations for tarsal pollen
collection in the Bombyliidae (Insecta: Diptera):the benefits
of putting your foot in your mouth. Zootaxa 157: 1–14.
Neville, E. M. 1985. The epidemiology of Parafilaria bovicola
in the Transvaal Bushveld of South Africa. Onderstepoort
Journal of Veterinary Research 52: 261–267.
Newton, G. L., C. V. Booram, R. W. Barker, and O. M. Hale.
1977. Dried Hermetia illucens larvae meal as a supplement
for swine. Journal of Animal Science 44: 395–400.
Nielsen, P., O. Ringdahl, and S. L. Tuxen. 1954. 48a. Diptera 1
(exclusive of Ceratopogonidae and Chironomidae). Zoology
Iceland 3: 1–189.
Norrbom, A. L. 2004. Fruit Fly (Diptera: Tephritidae)
Taxonomy Pages. http://www.sel.barc.usda.gov/diptera/
tephriti/tephriti.htm [Accessed 10 February 2008].
Noutis, C. and L. E. Millikan. 1994. Myiasis. Dermatologic
Clinics 12: 729–736.
O’Grady, P., J. Bonacum, R. DeSalle, and F. Do Val. 2003.
The placement of Engiscaptomyza, Grimshawomyia,and
Titanochaeta, three clades of endemic Hawaiian Drosophili-
dae (Diptera). Zootaxa 159: 1–16.
O’Hara, J. E., K. D. Floate, and B. E. Cooper. 1999. The Sar-
cophagidae (Diptera) of cattle feedlots in southern Alberta.
Journal of the Kansas Entomological Society 72: 167
176.
Økland, B. 1994. Mycetophilidae (Diptera), an insect group
vulnerable to forest practices? A comparison of clearcut,
managed and semi-natural spruce forests in southern
Norway. Biodiversity and Conservation 3: 68–85.
Økland, B. 1996. Unlogged forests: Important sites for pre-
serving the diversity of mycetophilids (Diptera, Sciaroidea).
Biological Conservation 76: 297–310.
Økland, B. 2000. Management effects on the decomposer
fauna of Diptera in spruce forests. Studia Dipterologica 7:
213–223.
Økland, B., F. G ¨otmark, B. Nord´en, N.Franc, O.Kurina, and
A. Polevoi 2004. Regional diversity of mycetophilids
(Diptera: Sciaroidea) in Scandinavian oak-dominated
forests. Biological Conservation 121: 9–20.
Økland, B., F. G ¨otmark, and B. Nord ´en. 2008. Oak wood-
land restoration: testing the effects on biodiversity of
mycetophilids in southern Sweden. Biodiversity and Con-
servation, 17 (11): 2599–2616.
Oldroyd, H. 1964. The Natural History of Flies. Weidenfield and
Nicholson, London. xiv +324 pp.
218 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
Oldroyd, H. 1973. Tabanidae (horse-flies, clegs, deer-flies,
etc.). Pp. 195– 202. In K. G. V. Smith (ed). Insects and Other
Arthropods of Medical Importance. British Museum (Natural
History), London. xiv +561 pp.
O’Meara, G. F. 1976. Saltmarsh mosquitoes (Diptera: Culici-
dae). Pp. 303–333. In L. Cheng (ed). Marine Insects.North-
Holland Publishing, Amsterdam, The Netherlands. 581 pp.
Oosterbroek, P. and G. W. Courtney. 1995. Phylogeny of the
Nematocerous families of Diptera (Insecta). Zoological Jour-
nal of the Linnean Society 115: 267–311.
Opit, G. P., B. Roitberg, and D. R. Gillespie. 1997. The func-
tional response and prey preference of Feltiella acarisuga
(Vallot) (Diptera: Cecidomyiidae) for two of its prey: male
and female twospotted spider mites, Tetranychus urticae
Koch (Acari: Tetranychiidae). Canadian Entomologist 129:
221–227.
Pagnier, J., J. G. Mears, O. Dunda-Belkhodja, K. E. Schaefer-
Rego, C. Beldjord, R. L. Nagel, and D. Labie. 1984. Evidence
for the multicentric origin of the sickle cell hemoglobingene
in Africa. Proceedings of the National Academy of Sciences USA
81: 1771–1773.
Palmeri, V., S. Longo, and A. E. Carolei. 2003. Osservazioni
su ‘‘Apimiasi’’ causate da Senotainia tricuspis in Calabria.
Apicoltore Moderno 88: 183–189.
Panzer, G. W. F. 1794. Fauna insectorvm germanicae initia oder
Deutschlands Insecten. Heft 24. Felsecker, N ¨urnberg. 24 pp.,
24 pls.
Pape, T. 1996. Catalogue of the Sarcophagidae of the world
(Insecta: Diptera). Memoirs of Entomology International 8:
1–558.
Pape, T. 2009. 4. Economic importance of Diptera. In
B. V. Brown, A. Borkent, J. M. Cumming, D. M. Wood, and
M. Zumbado (eds). Manual of Central American Diptera. Vol-
ume 1. NRC Research Press, Ottawa.
Papp, L. 1976. Ecological and zoogeographical data on flies
developing in excrement droppings (Diptera). Acta zoologica
Hungarica 22: 119–138.
Papp, L. 1985. Flies (Diptera) developing in sheep droppings
in Hungary. Acta zoologica Hungarica 31: 367–379.
Papp, L., and B. Darvas (eds). 2000. Contributions to a Manual of
Palaearctic Diptera. Volume 1. General and Applied Dipterology.
Science Herald, Budapest. 978 pp.
Papp, L. and P. Garz ´o. 1985. Flies (Diptera) of pasturing cat-
tle: some new data and new aspects. Folia entomologica
Hungarica 46: 153–168.
Parrella, M. P. 2004. UC IPM pest management guide-
lines: floriculture and ornamental nurseries. UC ANR
Publication 3392. http://www.ipm.ucdavis.edu/PMG/
r280301111.html [Accessed 15 October 2004].
Parsonson, I. M. 1993. Bluetongue virus infection of cattle.
Pp. 120–125. In Proceedings of the 97th Annual Meeting of
the USAHA. Cummings Corporation and Carter Printing
Company, Richmond. 577 pp.
Peacock, L. and G. A. Norton. 1990. A critical analysis of
organic vegetable crop protection in the U.K. Agriculture,
Ecosystems and Environment 31: 187–197.
Pellmyr, O.1989. The cost of mutuali sm: interactions between
Trollius europaeusand its pollinating parasites. Oecologia 78:
53–59.
Pitkin, B. R. 1989. Family Cryptochetidae. P. 666. In
N. L. Evenhuis (ed). Catalog of the Diptera of the Australasian
and Oceanian Regions. Bishop Museum Special Publication
86. Bishop Museum Press, Honolulu, and E.J. Brill, Leiden.
1155 pp.
Platt, A. P. and S. J. Harrison. 1995. Robber fly and trout pre-
dation on adult dragonflies (Anisoptera: Aeshnidae) and
first records of Aeshna umbrosa from Wyoming. Entomologi-
cal News 106: 229–235.
Pollard, G. 2000. FAO Assistance to Jamaica for improving
hot pepper production, in part through development of an
IPM programme for gall midges and broad mite. Caribbean
IPM Knowledge Network, Current Awareness Bulletin 2:
1–6.
Pollet, M. 1992. Impact of enviromental variables on the
occurrence of dolichopodid flies in marshland habitats in
Belgium (Diptera: Dolichopodidae). Journal of Natural His-
tory 26: 621–636.
Pollet, M. 2000. A documented red list of the dolichopodid
flies (Diptera: Dolichopodidae) of Flanders. Communications
of the Institute of Nature Conservation 8: 1–190. [In Dutch,
with English summary.]
Pollet, M. 2001. Dolichopodid biodiversity and site quality
assessment of reed marshes and grasslands in Belgium
(Diptera: Dolichopodidae). Journal of Insect Conservation 5:
99–116.
Pollet, M. and P. Grootaert 1996. An estimation of the nat-
ural value of dune habitats using Empidoidea (Diptera).
Biodiversity and Conservation 5: 859–880.
Pont, A. C. 1980. Family Mormotomyiidae. P. 713. In
R. W. Crosskey (ed). Catalogue of the Diptera of the Afro-
tropical Region. British Museum (Natural History), London.
1437 pp.
Porter, S. D., L. A. Nogueirade S ´a, and L. W. Morrison. 2004.
Establishment and dispersal of the fire ant decapitating fly
Pseudacteon tricuspis in north Florida. Biological Control 29:
179–188.
Potts, W. H. 1973. Glossinidae (tse-tse flies). Pp. 209–249. In
K. G. V. Smith (ed). Insects and Other Arthropods of Medical
Importance. British Museum (Natural History), London. xiv
+561 pp.
Powers, N. R. and S. E. Cope. 2000. William Crawford Gorgas:
the great sanitarian. Wing Beats 11: 4–28.
Pritchard, G. 1983. Biology of Tipulidae. Annual Review of
Entomology 28: 1–22.
Raddum, G. G. and O. A. Saether. 1981. Chironomid com-
munities in Norwegian lakes with different degrees of
acidification. Verhandlungen der Internationalen Vereinigung
f¨
ur Theoretische und Angewandte Limnologie. 21: 399–
405.
Ram´
ırez, W. 1984. Biolog´
ıa del g´enero Melaloncha (Phoridae),
moscas parasitoides de la abeja dom ´estica (Apis mellifera L.)
en Costa Rica. Revista de Biologia Tropical 32: 25–28.
Biodiversity of Diptera 219
Reemer, M. 2005. Saproxylic hoverflies benefit by modern
forest management (Diptera: Syrphidae). Journal of Insect
Conservation 9: 49–59.
Reiter, L. T., L. Potocki, S. Chien, M. Gribskov, and E. Bier.
2001. A systematic analysis of human disease-associated
gene sequences in Drosophila melanogaster.Genome Research
11: 1114–1125.
Resh, V. H., M. J. Myers, and M. J. Hannaford. 1996. Use of
biotic indices, habitat assessments, and benthicmacroinver-
tebrates in evaluating environmental quali ty. Pp. 647–667.
In F. R. Hauer and G. A. Lamberti (eds). Methods in Stream
Ecology. Academic Press, San Diego, California. 674 pp.
Reyes, F. 1983. A new record of Pseudohypocera kerteszi,apest
of honey bees in Mexico. American Bee Journal 119–120.
Ridsdill-Smith, T. J. 1981. Some effects of three species of
dung beetles (Coleoptera: Scarabaeidae) in south-western
Australia on the survival of the bush fly, Musca vetustis-
sima Walker (Diptera: Muscidae), in dung pads. Bulletin of
Entomological Research 71: 425–433.
Ridsdill-Smith, T. J., L. Hayles, and M. Palmer. 1987. Mortal-
ity of eggs and larvae of the bush fly, Musca vetustissima
Walker (Diptera: Muscidae) caused by scarabaeine dung
beetles (Coleoptera: Scarabaeidae) in favourable cattle
dung. Bulletin of Entomological Research 77: 731– 736.
Rinker, D. L. and R. J. Snetsinger. 1984. Damage threshold to
a commercial mushroom by a mushroom-infesting phorid
(Diptera: Phoridae). Journal of Economic Entomology 77:
449–453.
Robineau-Desvoidy, J. B. 1830. Essai sur les myodaires.
M´
emoires pr´
esent´
es par divers Savans a l’ Acad´
emie Royal
des Sciences de l’ Institut de France 2 (2): 1–813.
Robinson, G. E. 1981. Pseudohypocera kerteszi (Enderlein)
(Diptera: Phoridae), a pest of the honey bee. Florida Ento-
mologist 64: 456–457.
Robles, C. D. and J. Cubit. 1981. Influence of biotic factors in
an upper intertidal community: dipteran larvae grazing on
algae. Ecology 62: 1536–1547.
Robroek, B. J. M., H. de Jong, H. G. Arce, and M. J. Sommeijer.
2003a. The development of Pseudohypocera kerteszi (Diptera,
Phoridae), a kleptoparasite in nests of stingless bees
(Hymenoptera, Apidae) in Central America. Proceedings of
the Section Experimental and Applied Entomology, Nederlandse
Entomologische Vereniging, Amsterdam 12: 71–74.
Robroek, B. J. M., H. de Jong, and M. J. Sommeijer. 2003b.
The behaviour of Pseudohypocera kerteszi (Diptera, Phori-
dae) in hives of stingless bees (Hymenoptera, Apidae) in
Central America. Proceedings of the Section Experimental and
Applied Entomology, Nederlandse Entomologische Vereniging,
Amsterdam 12: 65–70.
Rogers, R. and M. Mattoni. 1993. Observations on the natural
history and conservation biology of the giant flower lov-
ing flies, Rhaphiomidas (Diptera: Apioceridae). Dipterological
Research 4 (1–2): 21 –34.
Rosenberg, D. M. and V. H. Resh (eds). 1993. Freshwater
Biomonitoring and Benthic Macroinvertebrates. Chapman and
Hall, New York. 488 pp.
Rosenberg, D. M., V. H. Resh, and R. S. King. 2007. Use of
aquatic insects in biomonitoring. Pp. 123–137. In R. W.
Merritt, K. W. Cummins, and M. B. Berg (eds). An Introduc-
tion to the Aquatic Insects of North America. Fourth Edition.
Kendall/Hunt Publishing, Dubuque, Iowa. 1158 pp.
Ross, A. 1961. Biological studies on bat ectoparasites of the
genus Trichobius (Diptera: Streblidae) in North America,
north of Mexico. Wasmann Journal of Biology 19: 229– 246.
Rossi, M. A. and S. Zucoloto. 1973. Fatal cerebral myiasis
caused by the tropical warble fly, Dermatobia hominis.Amer-
ican Journal of Tropical Medicine and Hygiene 22: 267–269.
Rotheray, G. E. and I. MacGowan. 2000. Status and breeding
sites of three presumed endangered Scottish saproxylic
syrphids (Diptera, Syrphidae). Journal of Insect Conservation
4: 215–223.
Rotheray, G. E., G. Hancock, S. Hewitt, D. Horsfield,
I. MacGowan, D. Robertson, and K. Watt. 2001. The
biodiversity and conservation of saproxylic Diptera in
Scotland. Journal of Insect Conservation 5: 77–85.
Rozkoˇsn´y, R. 1997. Family Stratiomyidae. Pp. 387–411. In
L. Papp and B. Darvas (eds). Contributions to a Manual of
Palaearctic Diptera.Volume 2. Nematocera and Lower Brachyc-
era. Science Herald, Budapest. 592 pp.
Rubega, M. and C. Inouye. 1994. Prey switching in Red-
necked Phalaropes (Phalaropus lobatus): Feeding limitations,
the functional response and water management at Mono
Lake, California, USA.Biological Conservation 70: 205–210.
Sabrosky, C. W. 1987. Chloropidae. Pp. 1049–1067. In
J. F. McAlpine, B. V. Peterson, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). Manual of
Nearctic Diptera. Volume 2. Research Branch, Supply &
Services Canada, Hull, Quebec, Agricultural Canada Mono-
graph 28.
Sæther, O. A. 1979. Chironomid communities as water qual-
ity indicators. Holarctic Ecology 2: 65–74.
Sakai, S., M. Kato, and H. Nagamasu. 2000. Artocarpus
(Moraceae) – gall midge pollination mutualism mediated
by a male-flower parasitic fungus. American Journal of Botany
87: 440–445.
Sanchez-Arroyo, H. 2007. House fly. University of
Florida, Publication Number: EENY-48. http://creatures
.ifas.ufl.edu/urban/flies/house fly.htm [Accessed 10
February 2008.]
Santini, L. 1995a. Senotainia tricuspis (Meigen) (Diptera, Sar-
cophagidae) in littoral Tuscany apiaries. L’Ape nostra Amica
(1995): 4–10.
Santini, L. 1995b. On the ‘‘Apimyiasis’’ caused by Senotainia
tricuspis (Meigen) (Diptera, Sarcophagidae) in Central Italy.
Apicoltore Moderno 86: 179–183.
Sawabe, K., K. Hoshino, H. Isawa, T. Sasaki, T. Hayashi,
Y. Tsuda, H. Kurahashi,K. Tanabayashi, A. Hotta, T. Saito,
A. Yamada, and M. Kobayashi. 2006. Detection and isola-
tion of highly pathogenic H5N1 avian influenza A viruses
from blow flies collected in the vicinity of an infected poultry
farm in Kyoto, Japan, 2004. American Journal of Tropical
Medicine and Hygiene 75: 327–332.
220 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
Scheepmaker, J. W. A., F. P. Geels, P. H. Smits, and L. J. L. D.
Van Griensven. 1997. Location of immature stages of the
mushroom insect pest Megaselia halterata in mushroom-
growing medium. Entomologia Experimentalis et Applicata
83: 323–327.
Scholl, P. J. 1993. Biology and control of cattle grubs. Annual
Review of Entomology 39: 53–70.
Schumann, H. 1992. Systematische Gliederung der Ordnung
Diptera mit besonderer Ber ¨ucksichtigung der in Deutsch-
land vorkommenden Familien. Deutsche Entomologische
Zeitschrift 1–3: 103 –116.
Sheppard, D. C. 1983. House fly and lesser house fly control
utilizing the black soldier fly in manure management sys-
tems for caged laying hens. Environmental Entomology 12:
1439–1242.
Sheppard, D. C., G. L. Newton, and S. A. Thompson. 1994. A
value added manure management system using the black
soldier fly. Bioresource Technology 50: 275–279.
Sherman, R. A. 2001. Maggot therapy for foot and leg
wounds. International Journal of Lower Extremity Wounds
1: 135–142.
Sherman, R. A. 2002. Maggot vs conservative debridement
therapy for the treatment of pressure ulcers. Wound Repair
and Regeneration 10: 208–214.
Sherman, R. A. 2003. Cohort study of maggot therapy for
treating diabetic foot ulcers. Diabetes Care 26: 446–
451.
Sherman, R. A., M. J. R. Hall, and S. Thomas. 2000. Medici-
nal maggots: an ancient remedy for some contemporary
afflictions. Annual Review of Entomology 45: 55–81.
Sinclair, B. J. 1988. The madicolous Tipulidae (Diptera) of
eastern North America, with descriptions of the biology and
immature stages of Dactylolabis montana (O.S.) and D. hud-
sonica Alexander (Diptera: Tipulidae). Canadian Entomologist
120: 569–573.
Sinclair, B. J. 1989. The biology of Euparyphus Gerstaecker
and Caloparyphus James occurring in madicolous habitats
of eastern North America, with descriptions of adult and
immature stages (Diptera: Stratiomyiidae). Canadian Journal
of Zoology 67: 33–41.
Sinclair, B. J. 2000. Immature stages of Australian Aus-
trothaumalea Tonnoir and Niphta Theischinger (Diptera:
Thaumaleidae). Australian Journal of Entomology 39:
171–176.
Sinclair, B. J. and J. M. Cumming. 2006. The morphology,
higher-level phylogeny and classification of the Empidoidea
(Diptera). Zootaxa 1180: 1–172.
Sinclair, B. J. and S. A. Marshall. 1987. The madicolous fauna
of southern Ontario. Proceedings of the Entomological Society
of Ontario 117: 9–14.
Sivinski, J. 1997. Ornaments in the Diptera. Florida Entomolo-
gist 80: 142–164.
Skevington, J. H. 2000. Pipunculidae (Diptera) systematics:
spotlight on the diverse tribe Eudorylini in Australia. Ph. D.
thesis. University of Queensland, Brisbane. 446 pp.
Skevington, J. H. 2001. Revision of Australian Clistoabdomi-
nalis (Diptera: Pipunculidae). Invertebrate Taxonomy 15 (5):
695–761.
Skevington, J. H. 2008. Hilltopping. Pp. 1799–1807. In
J. L. Capinera (ed) Encyclopedia of Entomology. 4. Second
Edition. Springer.
Skevington, J. H. and P. T. Dang (eds). 2002. Exploring the
diversity of flies (Diptera). Biodiversity 3: 3–27.
Skidmore, P. 1985. The Biology of the Muscidae of the World.
Junk, Dordrecht. Series Entomologica29. xiv +550 pp.
Skidmore, P. 1991. Insects of the British cow-dung com-
munity. Field Studies Council. Occasional Publication 21:
1–160.
Skovmand, O. 2004. Bti: control of mosquitoes and blackflies.
Biocontrol Files (2004, December issue): 7.
Smith, K. G. V. 1969. Diptera. Family Lonchopteridae. Hand-
books for the Identification of British Insects 10 (2ai):
1–9.
Somchoudhury, A. K., T. K. Das, P. K. Sarkar, and A. B.
Mukherjee. 1988. Life history of three diptera nfl ies infesting
edible mushroom Pleurotus sajor-caju.Environment and
Ecology 6: 463–464.
Sommagio, D. 1999. Syrphidae: can they be used as
environmental bioindicators? Agriculture, Ecosystems and
Environment 74: 343–356.
Sotavalta, O. 1947. The flight-tone (wing-stroke frequency)
of insects. Acta Entomologica Fennica 4: 1–117.
Sotavalta, O. 1953. Recordings of high wing-stroke and tho-
racic vibration frequency in some midges. Biological Bulletin
of Woods Hole 104: 439–444.
Speight, M. C. D. 1986. Attitudes to insects and insect con-
servation. Pp. 369–385. Proceedings of the 3rd European
Congress of Entomology.
Spencer, K. A. 1973. Agromyzidae (Diptera) of economic
importance. Series Entomologica 9. xi +418 pp.
Spencer, K. A. 1990. Host Specialization in the World Agromyzi-
dae (Diptera). Kluwer Academic Publishers, Dordrecht. xii
+444 pp.
Spofford, M. G. and F. E. Kurczewski. 1990. Comparative
larvipositional behaviours and cleptoparasitic frequencies
of Nearctic species of Miltogrammini (Diptera: Sarcophagi-
dae). Journal of Natural History 24: 731–755.
Ssymank, A. and D. Doczkal. 1998. Rote Liste der
Schwebfliegen (Diptera: Syrphidae). Pp. 65–72. In M. Binot,
R. Bless, P. Boye, H. Gruttke, and P. Pretscher (eds). Rote
Liste gef¨
ahrdeter Tiere Deutschlands. Schriftenreihe f ¨ur Land-
schaftspflege und Naturschutz 55.
Ssymank, A., C. A. Kearns, T. Pape, and F. C. Thompson.
2008. Pollinating flies (Diptera): a major contribution to
plant diversity and agricultural production. Biodiversity 9:
86–89.
Staerkeby, M. 2001. Dead larvae of Cynomya mortuorum
(L.) (Diptera, Calliphoridae) as indicators of the post-mortem
interval a case history from Norway. Forensic Science
International 120: 77–78.
Biodiversity of Diptera 221
Stalker, H. D. 1956. On the evolution of parthenogenesis in
Lonchoptera (Diptera). Evolution 10: 345–359.
Stark, A. 1996. Besonderheiten der Dipterenfauna Sachsen-
Anhalts-eine Herausforderung f ¨ur den Natur- und
Umweltschutz. Berichte des Landesamtes f¨
ur Umweltschutz
Sachsen-Anhalt, Halle 1996 (21): 100–108.
Steck, G. 2005. Pea leaf miner, Liriomyza huidobrensis
(Diptera: Agromyzidae). http://www.doacs.state.fl.us/pi/
enpp/ento/peamin.html [Last updated March 2006,
accessed 13 April 2007].
Steyskal, G. C. 1987. Pyrgotidae. Pp. 813–816. In
J. F. McAlpine, B. V. Peterson, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). Manual of
Nearctic Diptera. Volume 2. Research Branch, Supply &
Services Canada, Hull, Quebec, Agricultural Canada Mono-
graph 28.
Stireman, J. O., J. E. O’Hara, and D. M. Wood. 2006.
Tachinidae: evolution, behavior, and ecology. Annual
Review of Entomology 51: 525–555.
Struwe, I. 2005. Rhythmic dancing as courtship behaviour
in Limnophora riparia (Fall´en) (Diptera: Muscidae). Studia
dipterologica 11: 597–600.
Syme, D. A. and R. K. Josephson. 2002. How to build fast mus-
cles: synchronous and asynchronous designs. Integrative
and Comparative Biology 42: 762–770.
Sze, W. T., T. Pape, and D. K. O’Toole. 2008. The first blow fly
parasitoid takes a head start in its termite host (Diptera:
Calliphoridae, Bengaliinae; Isoptera: Macrotermitidae).
Systematics and Biodiversity 6: 25–30.
Szwejda, J. and R. Wrzodak. 2007. Phytophagous entomo-
fauna occurring on carrot and plant protection methods.
Vegetable Crops Research Bulletin 67: 95–102.
Taylor, G. S. 2004. Revision of Fergusonina Malloch gall
flies (Diptera: Fergusoninidae) from Melaleuca (Myrtaceae).
Invertebrate Systematics 18: 251–290.
Teskey, H. J. 1976. Diptera larvae associated with trees in
North America. Memoirs of the Entomological Society of
Canada 100: 1–53.
Teskey, H. J. 1981. Vermileonidae. Pp. 529–532. In
J. F. McAlpine, B. V. Petersen, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). Manual of
Nearctic Diptera. Volume 1. Research Branch, Supply &
Services Canada, Hull, Quebec, Agriculture Canada Mono-
graph 27.
Thompson, F. C. and G. E. Rotheray. 1998. Family Syrphidae.
Pp. 81–139. In L. Papp and B. Darvas (eds). Contributions to
a Manual of Palaearctic Diptera. Volume 3. General and Applied
Dipterology. Science Herald, Budapest. 880 pp.
Thomson, J. R., D. D. Hart, D. F. Charles, T. L. Nightengale,
and D. M. Winter. 2005. Effects of removal of a small dam
on downstream macroinvertebrate and algal assemblages
in a Pennsylvania stream. Journal of the North American
Benthological Society 24: 192–207.
Tomberlin, J. K., W. K. Reeves, and D. C. Sheppard. 2001.
First record of Chrysomya megacephala (Diptera:
Calliphoridae) in Georiga [sic], U.S.A.Florida Entomologist
84: 300–301.
Tondella, M. L., C. H. Paganelli, I. M. Bortolotto, O. A. Takano,
K. Irino, M. C. Brandileone, B. Mezzacapa Neto, V. S. Vieira,
and B. A. Perkins. 1994. Isolation of Haemophilus aegyptius
associated with Brazilian purpuric fever, of Chloropidae
(Diptera) of the genera Hippelates and Liohippelates.Revista
do Instituto de Medicina Tropical de S˜
ao Paulo 36: 105–109.
Turner, C. E. 1996. Tephritidae in the biological control of
weeds. Pp. 157–164. In B. A. McPheron and G. J. Steck
(eds). Fruit Fly Pests: A World Assessment of Their Biology
and Management. St. Lucie Press, Delray Beach. xxii +
586 pp.
USDA, Forest Service 1985. Insects of Eastern Forests. Miscel-
laneous Publications 1426: 1–608.
Vail, P. V., J. R. Coulson, W. C. Kauffmann, and M. E. Dix.
2001. History of biological control programs in the United
States Department of Agriculture. American Entomologist
47: 24–49.
Vaillant, F. 1956. Recherches sur la faune madicole
(hygrop´etrique s.l.) de France, de Corse et d’ Afrique de
Nord. M´
emoirs du Museum d’Histoire Naturelle Paris 11:
1–258.
Vaillant, F. 1961. Fluctuations d’une population madicole
au cours d’une ann´ee. Verhandlungen der Internationalen
Vereinigung fur Theoretische und Angewandte Limnologie 14:
513–516.
Val, F. C. 1992. Pantophthalmidae of Central America and
Panama (Diptera). Pp. 600–610. In D. Quintero and
A. Aiello (eds). Insects of Panama and Mesoamerica: Selected
Studies. Oxford University Press, Oxford. 692 pp.
Valentin, A., M. P. Baumann, E. Schein, and S. Bajanbileg.
1997. Genital myiasis (Wohlfahrtiosis) in camel herds of
Mongolia. Veterinary Parasitology 73: 335–346.
Van Riper C., III, S. G. Van Riper, and W. R. Hansen. 2002.
Epizootiology and effect of avian pox on Hawaiian forest
birds. Auk 119: 949–942.
Vigoreaux, J. O. 2001. Gene tics of the Drosophilaflight muscle
myofibril: a window into the biology of complex systems.
Bioessays 23: 1047–1063.
Vockeroth, J. R. 2002. Introducing the ubiquitous Diptera.
Biodiversity 3(4):35.
Wall, R. and D. Shearer. 1997. Veterinary Entomology.
Chapman and Hall, London. 439 pp.
Wall, R., J. J. Howard, and J. Bindu. 2001. The seas onal abun-
dance of blowflies infesting drying fish in south-west India.
Journal of Applied Ecology 38: 339–348.
Ward, J. V. 1992. Aquatic Insect Ecology. 1.Biology and
Habitat. John Wiley and Sons, Inc., New York. 438 pp.
Ward, J. V. 1994. Ecology of alpine streams. Freshwater Biol-
ogy 32: 277–294.
Warwick, W. F. 1989. Morphological deformities in larvae of
Procladius Skuse (Diptera: Chironomidae) and their biomo n-
itoring potential. Canadian Journal of Fisheries and Aquatic
Science 46: 1255–1270.
222 Gregory W. Courtney, Thomas Pape, Jeffrey H. Skevington, et al.
Warwick, W. F. 1991. Indexing deformities in ligulae and
antennae of Procladius larvae (Diptera: Chironomidae):
application to contaminant-stressed environments. Cana-
dian Journal of Fisheries and Aquatic Science 48: 1151 –1166.
Warwick, W. F. and N. A. Tisdale. 1988. Morphologica l defor-
mities in Chironomus, Cryptochironomus and Procladius lar-
vae (Diptera: Chironomidae) from two differentially stressed
sites in Tobin Lake, Saskatchewan. Canadian Journal of Fish-
eries and Aquatic Science 45: 1123– 1144.
Wasmann, E. 1910. Modern Biology and the Theory of Evo-
lution. Herder, St. Louis, MO. 539 pp., 8 pls. [Translated
by A. M. Buchanan from E. Wasmann. 1906. Die moderne
Biologie und die Entwicklungstheorie, Third Edition. Freiburg.
528 pp.].
Waterhouse, D. F. and D. P. A. Sands. 2001. Classical biolog-
ical control of arthropods in Australia. ACIAR Monograph
Series 77: 1–560.
Wells, J. D. 1991. Chrysomya megacephala (Diptera: Calliphori-
dae) has reached the continental United States: review of its
biology, pest status, and spread around the world. Journalof
Medical Entomology 28: 471–473.
Westwood, J. O. 1852. Observations on the destructive species
of dipterous insects known in Africa under the names of
tsetse, zimb and tsaltsalya, and on their supposed connec-
tion with the fourth plague of Egypt. Proceedings of the
Zoological Society of London 18: 258– 270.
Wheeler, W. M. 1930. Demons of the Dust.W.W. Norton and
Company, New York. xviii +378 pp.
White, I. M. and S. L. Clement. 1987. Systematic notes on
Urophora (Diptera, Tephritidae) species associated with Cen-
taurea solstitialis (Asteraceae, Cardueae) and other palaearc-
tic weeds adventive in North America. Proceedings of the
Entomological Society of Washington 89: 571–580.
White, P. F. 1981. Spread of the mushroom disease Verticil-
lium fungicola by Megaselia halterata (Diptera: Phoridae).
Protection Ecology 3: 17–24.
White, I. M. and M. M. Elson-Harris. 1992. Fruit Flies of Eco-
nomic Significance: Their Identification and Bionomics.CAB
International, Wallingford. xii +601 pp.
Whiteman, N. K., S. J. Goodman, B. J. Sinclair, T. Walsh,
A. A. Cunningham, L. D. Kramer, and P. G. Palmer. 2005.
Establishment of the avian disease vector Culex quinque-
fasciatus Say 1823 (Diptera: Culicidae) on the Gal ´apagos
Islands, Ecuador. Ibis 147: 844–847.
Wiederholm, T. 1980. Chironomids as indicators of water
quality in Swedish lakes. Acta Universitatis Carolinae-
Biologica 1978: 275–283.
Wiederholm, T. 1984. Incidence of deformed chironomid lar-
vae (Diptera: Chironomidae) in Swedish lakes. Hydrobiologia
109: 243–249.
Wiegmann, B. M., S.-C. Tsaur, D. W. Webb, D. K. Yeates, and
B. K. Cassel. 2000. Monophyly and relationship of the
Tabanomorpha (Diptera: Brachycera) based on 28S
ribosomal gene sequences. Annals of the Entomological Soci-
ety of America 93: 1031–1038.
Wood, D. M. 1981. Axymyiidae. Pp. 209–212. In
J. F. McAlpine, B. V. Petersen, G. E. Shewell, H. J. Teskey,
J. R. Vockeroth, and D. M. Wood (coordinators). Manual of
Nearctic Diptera. Volume 1. Research Branch, Supply &
Services Canada, Hull, Quebec, Agriculture Canada Mono-
graph 27.
Wood, D. M. and A. Borkent. 1989. Phylogeny and classifica-
tion of the Nematocera. Pp. 1333–1370. In J. F. McAlpine
and D. M. Wood (eds). Manual of Nearctic Diptera Volume 3.
Research Branch, Supply & Services Canada, Hull, Quebec,
Agricultural Canada Monograph 32.
Wood, D. M., P. T. Dang, and R. A. Ellis. 1979. The Mosquitoes
of Canada. Diptera: Culicidae. The Insects and Arachnids
of Canada, Part 6. Agriculture Canada, Hull, Quebec.
390 pp.
Woodley, N. E. and D. J. Hilburn. 1994. The Diptera of
Bermuda. Contributions of the American Entomological Insti-
tute 28: 1–64.
Wooton, R. 2000. From insects to microvehicles. Nature 403:
144–145.
World Bank. 1993. World Development Report 1993 –
Investing in Health. Oxford University Press, Oxford. 344 pp.
World Health Organization. 2007. Malaria fact sheet.
http://www.who.int/mediacentre/factsheets/fs094/en/
index.html [Accessed 10 February 2008].
Yeates, D. K. 1994. Cladistics and classifica tion of the Bombyli-
idae (Diptera: Asiloidea). Bulletin of the American Museum of
Natural History 219: 1–191.
Yeates, D. K. and B. M. Wiegmann. 1999. Congruence and
controversy: toward a higher-level classification of Diptera.
Annual Review of Entomology 44: 397–428.
Yeates, D. K., B. M. Wiegmann, G. W. Courtney, R. Meier,
C. Lambkin, and T. Pape. 2007. Phylogeny and system-
atics of Diptera: two decades of progress and prospects.
Zootaxa 1668: 565–590.
Yoder, C. O. and E. T. Rankin. 1998. The role of biolog-
ical indicators in a state water quality management
process. Environmental Monitoring and Assessment 51:
61–88.
Young, A. M. 1986. Cocoa pollination.Cocoa Growers Bulletin
37: 5.
Young, A. M. 1994. The Chocolate Tree, a Natural History of
Cacao. Smithsonian Institution Press, Washington,London.
200 pp.
Zhang, J. F. 1987. Four new genera of Platypezidae. Acta
Palaeontologica Sinica 26: 595–603.
Zumpt, F. 1965. Myiasis in Man and Animals in the Old World.
Butterworth, London. xv +267 pp.
Zwick, P. 1977. Australian Blephariceridae (Diptera).
Australian Journal of Zoology, Supplementary Series 46:
1–121.
... As insects that undergo holometabolous development, all aquatic Diptera have a life cycle that includes a series of different stages or instars. A typical fly life cycle consists of a short-term egg stage (usually a few days or weeks, only sometimes much longer), three or four larval instars (rarely more), a pupal stage of variable length, and an adult stage lasting from fewer than two hours (family Deuterophlebiidae) to several weeks or months (Courtney et al. 2017;Lackmann and Butler 2018;Adler and Courtney 2019). ...
... Utmost of aquatic Diptera are free-living insects that need a wet environment in at least one part of their life cycle (Adler and Courtney 2019) or, in more strict sense, aquatic Diptera are considered as those linked with water bodies (Courtney et al. 2017). Until now, we recognize 158 dipteran families worldwide, but only 41 families have aquatic representatives (Adler and Courtney 2019), and in Europe there are only 130 families of Diptera, from which about 25 of them are connected to aquatic habitats (Oosterbroek 2006). ...
Chapter
Aquatic insects are the most species-rich group that inhabit freshwaters. They are connected to water by at least one life stage, usually that of the larvae, and some spend their entire life in freshwater habitats. The majority of aquatic insects’ larvae develop in water; while adults emerge and spend their lives primarily in terrestrial environments where they mate, disperse and in some cases feed. Aquatic insects are sensitive to environmental conditions, which is why they are wiedly used in biomonitoring. Mayflies, stoneflies and caddisflies are among the most commonly used indicators, but other taxa also show high potential. In the Plitvice Lakes, there are confirmed records of 352 species of aquatic insects. The most abundant order of aquatic insects is Diptera, with 165 recorded species, followed by Trichoptera with 91 species, and Plecoptera with 31 species. All other aquatic insect orders are present with lower species richness, with 25 taxa recorded in Coleoptera, 18 taxa in Ephemeroptera and 14 in Odonata. Megaloptera and Neuroptera each have four species. The aquatic insects reflect the uniqueness and peculiarities of the barrage system of the Plitvice Lakes in both, their composition and ecology. Although some groups of aquatic insects have been well studied throughout the years, there are still many unknowns and many species that need to be recorded.KeywordsAquatic insectsEphemeropteraOdonataPlecopteraMegalopteraNeuropteraColeopteraDipteraTrichopteraSpecies richnessEmergence traps
... On a global scale, there is a great diversity and wide distribution of Diptera (Arthropoda, Insecta) species. The existence of more than 157 000 species, 10 000 genera, and 188 families of flies, horseflies, and mosquitoes has been estimated (Courtney et al. 2017). ...
... The diversity of Diptera associated with the milking environment recorded in this study results from the variety of eating and behavioral habits that have been reported for this taxon (Courtney et al. 2017). From the identification of a considerable portion of the species, it was possible to observe that the organic matter, mainly feces and manure present in the farm areas, and the domestic animals themselves have a relevant role in the attractiveness of these dipterans in search of energy resources (Rafael et al. 2009) since more than 90% of the species were coprophagous (M. ...
Article
Full-text available
Aims: We aimed to investigate the prevalence of rotavirus and coronavirus in dipterans that commonly inhabit the environment of dairy farms. Methods and results: We collected 217 insect specimens from nine dairy farms, which were examined through hemi-nested RT-PCR followed by Sanger sequencing in search of VP1 and N genes for rotavirus and bovine coronavirus-BCoV, respectively. Eleven families of Diptera were identified, with a predominance of Muscidae (152/217=70%). Rotavirus A (RVA) and BCoV were detected in 14.7% (32/217) and 4.6% (10/217) of the dipterans, respectively. Sequencing of the amplicons was possible for 11.5% (25/217) of RVA and 0.5% (1/217) of BCoV, confirming the presence of these pathogens. Conclusions: Our findings highlight the role of dipterans as carriers of rotavirus A and betacoronavirus of great relevance for public and animal health.
... Platystomatidae is a family of flies in the superfamily Tephritoidea (Steyskal 1961;Korneyev 1999). Currently, there are about 1,200 species in 120 genera (McAlpine 2001;Courtney et. al. 2009;Pape et al. 2011). The platystomatid fauna of the Americas comprises four genera, and 64 valid species: Amphicnephes Loew (three species); Himeroessa Loew (monotypic); Rivellia Robineau-Desvoidy (39 species), and Senopterina Macquart (21 species). Platystoma seminationis (Fabricius, 1775), Pogonortalis doclea (Wiedemann, 1819), and Plagi ...
Article
A review of Senopterina Macquart, 1835, a Nearctic and Neotropical genus with 21 species, is presented. Senopterina brevipes (Fabricius, 1805), the type species of the genus, and S. macularis (Fabricius, 1805), are redescribed. Two new species are listed for Brazil: S. canina sp. nov., and S. flavifemoris sp. nov., from the following states: Amapá, Goiás, Maranhão, Mato Grosso, Mato Grosso do Sul, Paraná, Rondônia and Tocantins. SEM images of the male and female terminalia of the redescribed species and of the new species are presented. An updated identification key to species is also provided.
... One notable family within this infraorder is Tabanidae, commonly known as horse or deer flies, which are known for their relatively large size and wide distribution. Nevertheless, there are smaller clades within the infraorder, with limited distributions, that are represented by relatively few species, such as Pelecorhynchidae and Austroleptidae (Courtney et al. 2009;Fachin et al. 2018Fachin et al. , 2020. ...
Article
The phylogeny of the genus Pelecorhynchus Macquart (Diptera: Pelecorhynchidae) was analyzed using three genes, cytochrome oxidase I, 28S ribosomal DNA, and CAD5, with 112 morphological characteristics. A total of 59 specimens (13 outgroups and 46 ingroups) were included in the analysis. The monophyly of Pelecorhynchidae was recovered under all analyses, with Glutops Burgess as the sister group of Pelecorhynchus s.l., while Pseudoerinna jonesi (Cresson) was the sister group. Within “Pelecorhynchus” there are three main clades with unresolved affinities. Clade I was formed by P. personatus (Walker), P. vulpes (Macquart), P. penai Pechuman and P. kroeberi (Lindner), a well-supported clade. Clade II corresponds to the set of species of “Chilean Pelecorhynchus”, conformed to P. biguttatus (Philippi), P. toltensis Llanos & González, P. elegans (Philippi), P. xanthopleura (Philippi), P. hualqui Llanos & González, and P. longicauda (Bigot), a well-supported clade. Clade III is represented exclusively by P. fulvus Ricardo, which has an exclusively Australian distribution. The monophyly of P. fulvus and its nomenclature remain an open question, as only a single species of this taxon was included. Our study demonstrated that the concept of Pelecorhynchus should be revisited. Therefore, we restore Coenura Bigot, 1857 to generic status for part of the southern South American species of “Pelecorhynchus” conformed by the species C. biguttata, C. elegans, C. hualqui, C. longicauda, C. toltensis, and C. xanthopleura which are monophyletic, supported by molecular and morphological data, and consistent with a Chilean distribution.
... Documenting extant biodiversity is a crucial challenge for contemporary science (Costello et al. 2013). Flies (Diptera) are one of the four super-radiations of insects, which together account for the majority of animal life of the Earth (Wiegmann et al. 2011, Courtney et al. 2017. Current knowledge of global Diptera diversity is still far from comprehensive (Pape et al. 2009), and it is particularly sparse for schizophoran flies, the group that evolved during the last episode of rapid dipteran radiations in the Early Cenozoic (Wiegmann et al. 2011, Bayless et al. 2021. ...
Article
Current knowledge of the global diversity of schizophoran flies, the megadiverse group that evolved during the episode of rapid dipteran radiations in the Early Cenozoic, is far from satisfactory. A significant part of this radiation is flesh flies (Sarcophagidae), which include about 3000 known species. Miltogramma Meigen, the largest genus of the flesh fly subfamily Miltogramminae, includes 119 known species with a peak of diversity in the Middle East and Central Asia. The taxonomy of the genus is acknowledged to be challenging, partly due to logistical issues associated with the deposition of type specimens in less accessible Russian institutions. Six new distinctive species of Miltogramma collected during recent expeditions to Iran and Uzbekistan are described using methods of integrative taxonomy: Miltogramma manouchehrii Szpila & Pape, sp. nov., M. mikolajczyki Szpila & Pape, sp. nov., M. rognesi Szpila & Pape, sp. nov., M. soszynskii Szpila & Pape, sp. nov., M. thompsoni Szpila & Pape, sp. nov., and M. vervesi Szpila & Pape, sp. nov. Flies are characterized by secondary sexual dimorphism expressed in the form of species-specific chaetotaxy of the male fore tarsus, a character related to courtship behaviour. The reconnaissance surveys to both of the mentioned countries performed in the last decade brought information about a large number of potential new species of Miltogramma, despite the taxonomic work performed in this area in the past. Describing the majority of them requires extensive revisionary work, which involves accessing the type specimens and applying modern methods of species delimitation.
... Among the best-known dipterans are the first genetic model organism Drosophila melanogaster, the disease vector mosquitos Anopheles and Culex, and the ubiquitous house fly Musca domestica. However, there are approximately 160 000 documented extant species of flies and this is likely to be a great underestimate [1]. Owing in part to their diversity and adaptability, dipterans have evolved multiple different reproductive strategies, from the use of both female homo-and hetero-gamety for sex determination to the use of parthenogenetic forms of reproduction [2,3]. ...
Article
Full-text available
Parthenogenesis has been documented in almost every phylum of animals, and yet this phenomenon is largely understudied. It has particular importance in dipterans since some parthenogenetic species are also disease vectors and agricultural pests. Here, we present a catalogue of parthenogenetic dipterans, although it is likely that many more remain to be identified, and we discuss how their developmental biology and interactions with diverse environments may be linked to different types of parthenogenetic reproduction. We discuss how the advances in genetics and genomics have identified chromosomal loci associated with parthenogenesis. In particular, a polygenic cause of facultative parthenogenesis has been uncovered in Drosophila mercatorum, allowing the corresponding genetic variants to be tested for their ability to promote parthe-nogenesis in another species, Drosophila melanogaster. This study probably identifies just one of many routes that could be followed in the evolution of parthenogenesis. We attempt to account for why the phenomenon has evolved so many times in the dipteran order and why facultative parthenogenesis appears particularly prevalent. We also discuss the significance of coarse geno-mic changes, including non-disjunction, aneuploidy, and polyploidy and how, together with changes to specific genes, these might relate to both facultative and obligate parthenogenesis in dipterans and other parthenogenetic animals.
Article
Subfossil pine and oak tree trunks were excavated during exploitation of the Budwity peatland in Northern Poland. Based on dendrochronological analysis, the woodland successions in peatland were reconstructed and correlated with moisture dynamics of the peatland ecosystem inferred from the high-resolution multi-proxy analysis of the peatland deposits. From the results of dendrochronological analysis and the 14C wiggle matching methods, four floating pine chronologies (5882–5595; 5250–5089; 3702–3546; and 2222–1979 mod. cal BP) and two oak chronologies (4932–4599 and 4042–3726 mod. cal BP) were developed. The organic sediments of the peatland (6 m thick) were deposited over approximately nine thousand years. The lower complex (525–315 cm) comprises minerogenic peat, while the upper complex (315.0–0.0 cm) is composed of ombrogenic peat. Subfossil tree trunks are distributed across various peat horizons, which suggests multiple stages of tree colonisation followed by subsequent dying-off phases. Multiproxy sediment analyses (lithological, geochemical and δ13C stable isotope, pollen, plant macrofossils, Cladocera, diatom, and Diptera analyses) indicate that the two earliest phases of pine colonisation (5882–5595 and 5250–5089 mod. cal BP) and the two stages of oak colonisation (4932–4599 and 4042–3726 mod. cal BP) were associated with periodic drying of the peatland. Conversely, tree dying-off phases occurred during periods of increased water levels in the peatland, coinciding with stages of increasing climate humidity during the Holocene. The two most recent phases of pine colonisation occurred during the ombrogenic stage of mire development. Remnants of the dead forest from these phases, marked by subfossil trunks still rooted in the ground, were preserved and exposed presently during peat exploitation, approximately 2.5 m below ground level. The identified phases of tree colonisation and subsequent dying-off phases show correlation with analogical phenomena observed in the other investigated European peatlands.
Article
Full-text available
Resumen. Se entrega una actualización de la diversidad de especies, géneros y familias para los microlepidópteros presentes en Chile (desde Micropterigoidea a Zygaenoidea) y una estimación de la diversidad taxonómica para el orden Lepidoptera en el país, de acuerdo con la información publicada en la literatura. Palabras clave: Diversidad de insectos; fauna; Microlepidoptera; Obtectomera; polillas. Abstract. An update of the diversity of species, genera, and families of the Microlepidoptera present in Chile (from Micropterigoidea to Zygaenoidea) and an estimate of the taxonomic diversity for the order Lepidoptera in the country, are provided according to the information published in the literature.
Article
Blow flies are of medical, sanitary, veterinary, and forensic importance. Their accurate taxonomic identification is essential for their use in applied research. However, neotropical fauna has not been completely studied or described, and taxa identification without the required training is a difficult task. Additionally, the current morphological keys are not fitting to all extant taxa. Molecular-based approaches are widely used to overcome these issues, including the standard 5' COI barcode fragment (~650 base pairs [bp]) for identification at the species level. Here, a shorter sequence of 5' COI fragment (~342 bp) was assessed for the identification of 28 blow fly species inhabiting the northwest of South America. One tree-based (the generalized mixed Yule-coalescent-GMYC) and 3 distance-based approaches (automatic barcode gap discover - ABGD, the best close match - BCM, and the nearest neighbor - NN) analyses were performed. Noticeably, the amplification and sequencing of samples that had been preserved for up to 57 years were successful. The tree topology assigned 113 sequences to a specific taxon (70% effectiveness), while the distance approach assigned to 95 (59% effectiveness). The short fragment allowed the molecular identification of 19 species (60% of neotropical species except for the Lucilia species and Hemilucilia semidiaphana). According to these findings, the taxonomic and faunistic considerations of the blow fly fauna were provided. Overall, the short fragment approach constitutes an optimal species confirmation tool for the most common blow flies in northwestern South America.
Article
The Pantophthalmidae is a small family of Diptera found exclusively in Neotropical forests. These flies are relatively rare in collections; and very scanty materials have been added since the revision by Val (1976). The 20 species, accepted as valid, are placed in two genera: Opetiops, with one species, known only from South America, and Pantophthalmus, with 19 species (Val 1976), ranging from Mexico to Argentina. Eleven species of Pantophthalmus are known to occur in Central America. Very little is known about the biology of the Pantophthalmidae. Even for the best studied species P. pictus (Andrade 1930) and P. tabaninus (Greene and Urich 1931; Bartholomew and Lighton 1986) there is no agreement about the feeding habits of the imago and larvae. The larvae live in galleries in trunks of dead or living trees. Andrade (1930) believed that the larvae were xylophagous while Thorpe (1934) had the opinion that they fed on fermenting sap. Greene and Urich (1931) believed that adults do not feed, while Thorpe (1934) argued that because the adults have mouthparts and a well-developed digestive system, they probably do feed. First instar larvae of P. tabaninus use orifices that are already present in trunks of trees and excavate their galleries using these orifices as the starting point, while in P. pictus the larvae begin excavation of their galleries in the absence of existing openings in the trunks. The larvae dig horizontal galleries in the trunks of various types of trees. P. pictus is found in live trees (Fieberg 1906; Andrade 1930), and P. tabaninus in decaying or fallen trees, but never in healthy ones (Greene and Urich 1931).