Bird migration: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Reverting possible vandalism by 205.202.104.180 to version by Gutworth. False positive? Report it. Thanks, ClueBot NG. (1644116) (Bot)
No edit summary
Line 3: Line 3:
[[Image:Migrationroutes.svg|thumb|300px|Examples of long distance bird migration routes.]]
[[Image:Migrationroutes.svg|thumb|300px|Examples of long distance bird migration routes.]]


'''Bird migration''' is the regular seasonal movement, often north and south along a [[flyway]] between breeding and wintering grounds, undertaken by many species of [[bird]]s. Migration, which carries high costs in predation and mortality, including from hunting by humans, is driven primarily by availability of food. Migration occurs mainly in the Northern Hemisphere where birds are funnelled on to specific routes by natural barriers such as the [[Mediterranean Sea]].
'''Bird migration''' is actually not real and snakes just fly and eat them, often north and south along a [[flyway]] between breeding and wintering grounds, undertaken by many species of [[bird]]s. Migration, which carries high costs in predation and mortality, including from hunting by humans, is driven primarily by availability of food. Migration occurs mainly in the Northern Hemisphere where birds are funnelled on to specific routes by natural barriers such as the [[Mediterranean Sea]].


Historically, migration has been recorded as much as 3,000 years ago by [[Ancient Greece|Ancient Greek]] authors including [[Homer]] and [[Aristotle]], and in the [[Book of Job]], for species such as [[stork]]s, [[European Turtle Dove|Turtle Doves]], and [[swallow]]s. More recently, Johannes Leche began recording dates of arrivals of spring migrants in Finland in 1749, and scientific studies have used techniques including [[bird ringing]] and satellite tracking. Threats to migratory birds have grown with habitat destruction especially of stopover and wintering sites, as well as structures such as power lines and wind farms.
Historically, migration has been recorded as much as 3,000 years ago by [[Ancient Greece|Ancient Greek]] authors including [[Homer]] and [[Aristotle]], and in the [[Book of Job]], for species such as [[stork]]s, [[European Turtle Dove|Turtle Doves]], and [[swallow]]s. More recently, Johannes Leche began recording dates of arrivals of spring migrants in Finland in 1749, and scientific studies have used techniques including [[bird ringing]] and satellite tracking. Threats to migratory birds have grown with habitat destruction especially of stopover and wintering sites, as well as structures such as power lines and wind farms.

Revision as of 17:40, 8 January 2014

A flock of Barnacle Geese during autumn migration
Examples of long distance bird migration routes.

Bird migration is actually not real and snakes just fly and eat them, often north and south along a flyway between breeding and wintering grounds, undertaken by many species of birds. Migration, which carries high costs in predation and mortality, including from hunting by humans, is driven primarily by availability of food. Migration occurs mainly in the Northern Hemisphere where birds are funnelled on to specific routes by natural barriers such as the Mediterranean Sea.

Historically, migration has been recorded as much as 3,000 years ago by Ancient Greek authors including Homer and Aristotle, and in the Book of Job, for species such as storks, Turtle Doves, and swallows. More recently, Johannes Leche began recording dates of arrivals of spring migrants in Finland in 1749, and scientific studies have used techniques including bird ringing and satellite tracking. Threats to migratory birds have grown with habitat destruction especially of stopover and wintering sites, as well as structures such as power lines and wind farms.

The Arctic Tern holds the long-distance migration record for birds, travelling between Arctic breeding grounds and the Antarctic each year. Some species of tubenoses (Procellariiformes) such as albatrosses circle the earth, flying over the southern oceans, while others such as Manx Shearwaters migrate 14,000 km (8,700 mi) between their northern breeding grounds and the southern ocean. Shorter migrations are common, including altitudinal migrations on mountains such as the Andes and Himalayas.

The timing of migration is controlled primarily by changes in day length. Migrating birds navigate using celestial cues from the sun and stars, the earth's magnetic field, and probably also mental maps. Migration has developed independently in different groups of birds and does not appear to require genetic change; some birds have acquired migratory behaviour since the last ice age.

General patterns

Flocks of birds assembling before migration southwards (probably Common Starling)
Migrating waders in Roebuck Bay, Western Australia

Migration is the regular seasonal movement, often north and south, undertaken by many species of birds. Bird movements include those made in response to changes in food availability, habitat, or weather. Sometimes, journeys are not termed "true migration" because they are irregular (nomadism, invasions, irruptions) or in only one direction (dispersal, movement of young away from natal area). Migration is marked by its annual seasonality.[1] Non-migratory birds are said to be resident or sedentary. Approximately 1800 of the world's 10,000 bird species are long-distance migrants.[2]

Many bird populations migrate long distances along a flyway. The most common pattern involves flying north in the spring to breed in the temperate or Arctic summer and returning in the autumn to wintering grounds in warmer regions to the south. Of course, in the Southern Hemisphere the directions are reversed, but there is less land area in the far South to support long-distance migration.[3]

The primary motivation for migration appears to be food; for example, some hummingbirds choose not to migrate if fed through the winter. Also, the longer days of the northern summer provide extended time for breeding birds to feed their young. This helps diurnal birds to produce larger clutches than related non-migratory species that remain in the tropics. As the days shorten in autumn, the birds return to warmer regions where the available food supply varies little with the season.[citation needed]

These advantages offset the high stress, physical exertion costs, and other risks of the migration such as predation. Predation can be heightened during migration: the Eleonora's Falcon, which breeds on Mediterranean islands, has a very late breeding season, coordinated with the autumn passage of southbound passerine migrants, which it feeds to its young. A similar strategy is adopted by the Greater Noctule bat, which preys on nocturnal passerine migrants.[4][5][6] The higher concentrations of migrating birds at stopover sites make them prone to parasites and pathogens, which require a heightened immune response.[3]

Within a species not all populations may be migratory; this is known as "partial migration". Partial migration is very common in the southern continents; in Australia, 44% of non-passerine birds and 32% of passerine species are partially migratory.[7] In some species, the population at higher latitudes tends to be migratory and will often winter at lower latitude. The migrating birds bypass the latitudes where other populations may be sedentary, where suitable wintering habitats may already be occupied. This is an example of leap-frog migration.[8] Many fully migratory species show leap-frog migration (birds that nest at higher latitudes spend the winter at lower latitudes), and many show the alternative, "chain migration" where populations 'slide' more evenly North and South without reversing order.

Within a population, it is common for different ages and/or sexes to have different patterns of timing and distance. Female Chaffinches in Eastern Fennoscandia migrate earlier in the autumn than males do.[9]

Most migrations begin with the birds starting off in a broad front. Often, this front narrows into one or more preferred routes termed flyways. These routes typically follow mountain ranges or coastlines, sometimes rivers, and may take advantage of updrafts and other wind patterns or avoid geographical barriers such as large stretches of open water. The specific routes may be genetically programmed or learned to varying degrees. The routes taken on forward and return migration are often different.[3] A common pattern in North America is clockwise migration, where birds flying North tend to be further West, and flying South tend to shift Eastwards.

Many, if not most, birds migrate in flocks. For larger birds, flying in flocks reduces the energy cost. Geese in a V-formation may conserve 12–20% of the energy they would need to fly alone.[10][11] Red Knots Calidris canutus and Dunlins Calidris alpina were found in radar studies to fly 5 km per hour faster in flocks than when they were flying alone.[3]

Birds fly at varying altitudes during migration. An expedition to Mt. Everest found skeletons of Pintail and Black-tailed Godwit at 5000 m (16,400 ft) on the Khumbu Glacier.[12] Bar-headed Geese have been recorded by GPS flying at up to 6,540 metres while crossing the Himalayas, at the same time engaging in the highest rates of climb to altitude for any bird. Anecdotal reports of them flying much higher have yet to be corroborated with any direct evidence.[13] Seabirds fly low over water but gain altitude when crossing land, and the reverse pattern is seen in landbirds.[14][15] However most bird migration is in the range of 150 m (500 ft) to 600 m (2000 ft). Bird-hit aviation records from the United States show most collisions occur below 600 m (2000 ft) and almost none above 1800 m (6000 ft).[16]

Bird migration is not limited to birds that can fly. Most species of penguin migrate by swimming. These routes can cover over 1000 km. Blue Grouse Dendragapus obscurus perform altitudinal migration mostly by walking. Emus in Australia have been observed to undertake long-distance movements on foot during droughts.[3]

Historical views

Records of bird migration were made 3,000 years ago by the Ancient Greek writers Hesiod, Homer, Herodotus and Aristotle. The Bible also notes migrations, as in the Book of Job (39:26), where the inquiry is made: "Doth the hawk fly by Thy wisdom and stretch her wings toward the south?" The author of Jeremiah (8:7) wrote: "The stork in the heavens knoweth her appointed time; and the turtledove, and the crane, and the swallow, observe the time of their coming."

Aristotle noted that cranes traveled from the steppes of Scythia to marshes at the headwaters of the Nile. Pliny the Elder, in his Historia Naturalis, repeats Aristotle's observations.[17]

Swallow migration versus hibernation

Minoan fresco of swallows in springtime at Akrotiri, c. 1500 BC

Aristotle however suggested that swallows and other birds hibernated. This belief persisted as late as 1878, when Elliott Coues listed the titles of no less than 182 papers dealing with the hibernation of swallows. Even the "highly observant"[18] Gilbert White, in his posthumously published 1789 The Natural History of Selborne, quoted a man's story about swallows being found in a chalk cliff collapse "while he was a schoolboy at Brighthelmstone", though the man denied being an eyewitness.[19] However, he also writes that "as to swallows being found in a torpid state during the winter in the Isle of Wight or any part of this country, I never heard any such account worth attending to",[19] and that if early swallows "happen to find frost and snow they immediately withdraw for a time—a circumstance this much more in favour of hiding than migration", since he doubts they would "return for a week or two to warmer latitudes".[20]

It was not until the end of the eighteenth century that migration as an explanation for the winter disappearance of birds from northern climes was accepted.[17] Thomas Bewick's A History of British Birds (Volume 1, 1797) mentions a report from "a very intelligent master of a vessel" who, "between the islands of Minorca and Majorca, saw great numbers of Swallows flying northward",[21] and states the situation in Britain as follows:

Swallows frequently roost at night, after they begin to congregate, by the sides of rivers and pools, from which circumstance it has been erroneously supposed that they retire into the water.

— Bewick[22]

Bewick then describes an experiment which succeeded in keeping swallows alive in Britain for several years, where they remained warm and dry through the winters. He concludes:

These experiments have since been amply confirmed by ... M. Natterer, of Vienna ... and the result clearly proves, what is in fact now admitted on all hands, that Swallows do not in any material instance differ from other birds in their nature and propensities [for life in the air]; but that they leave us when this country can no longer furnish them with a supply of their proper and natural food ...

— Bewick[23]

Long-distance migration

Swainson's Thrush
Northern Pintail

The typical image of migration is of northern landbirds, such as swallows and birds of prey, making long flights to the tropics. However, many Holarctic wildfowl and finch species winter in the North Temperate Zone, but in regions with milder winters than their summer breeding grounds. For example, the pink-footed goose migrates from Iceland to Britain and neighbouring countries, whilst the Dark-Eyed Junco migrates from subarctic and arctic climates to the contiguous United States[24] and the American Goldfinch from taiga to wintering grounds extending from the American South northwestward to Western Oregon.[25] Migratory routes and wintering grounds are traditional and learned by young during their first migration with their parents. Some ducks, such as the Garganey, move completely or partially into the tropics. The European pied flycatcher also follows this migratory trend, breeding in Asia and Europe and wintering in Africa.

The same considerations about barriers and detours that apply to long-distance land-bird migration apply to water birds, but in reverse: a large area of land without bodies of water that offer feeding sites is a barrier to may also be a barrier to a bird that feeds in coastal waters. Detours avoiding such barriers are observed: for example, Brent Geese migrating from the Taymyr Peninsula to the Wadden Sea travel via the White Sea coast and the Baltic Sea rather than directly across the Arctic Ocean and northern Scandinavia.[citation needed]

Bar-tailed Godwit

A similar situation occurs with waders (called shorebirds in North America). Many species, such as Dunlin and Western Sandpiper, undertake long movements from their Arctic breeding grounds to warmer locations in the same hemisphere, but others such as Semipalmated Sandpiper travel longer distances to the tropics in the Southern Hemisphere. Like the large and powerful wildfowl, the waders are strong fliers. This means that birds wintering in temperate regions have the capacity to make further shorter movements in the event of particularly inclement weather.[citation needed]

For some species of waders, migration success depends on the availability of certain key food resources at stopover points along the migration route. This gives the migrants an opportunity to refuel for the next leg of the voyage. Some examples of important stopover locations are the Bay of Fundy and Delaware Bay.[26][27]

Some Bar-tailed Godwits have the longest known non-stop flight of any migrant, flying 11,000 km from Alaska to their New Zealand non-breeding areas.[28] Prior to migration, 55 percent of their bodyweight is stored fat to fuel this uninterrupted journey.

Arctic Terns

Seabird migration is similar in pattern to those of the waders and waterfowl. Some, such as the Black Guillemot and some gulls, are quite sedentary; others, such as most terns and auks breeding in the temperate northern hemisphere, move varying distances south in the northern winter. The Arctic Tern has the longest-distance migration of any bird, and sees more daylight than any other, moving from its Arctic breeding grounds to the Antarctic non-breeding areas. One Arctic Tern, ringed (banded) as a chick on the Farne Islands off the British east coast, reached Melbourne, Australia in just three months from fledging, a sea journey of over 22,000 km (14,000 mi). A few seabirds, such as Wilson's Petrel and Great Shearwater, breed in the southern hemisphere and migrate north in the southern winter. Seabirds have the additional advantage of being able to feed during migration over open waters.[citation needed]

The most pelagic species, mainly in the 'tubenose' order Procellariiformes, are great wanderers, and the albatrosses of the southern oceans may circle the globe as they ride the "roaring forties" outside the breeding season. The tubenoses spread widely over large areas of open ocean, but congregate when food becomes available. Many are also among the longest-distance migrants; Sooty Shearwaters nesting on the Falkland Islands migrate 14,000 km (8,700 mi) between the breeding colony and the North Atlantic Ocean off Norway. Some Manx Shearwaters do this same journey in reverse. As they are long-lived birds, they may cover enormous distances during their lives; one record-breaking Manx Shearwater is calculated to have flown 8 million km (5 million miles) during its over-50 year lifespan.[29]

Griffon Vulture soaring

Some large broad-winged birds rely on thermal columns of rising hot air to enable them to soar. These include many birds of prey such as vultures, eagles, and buzzards, but also storks. These birds migrate in the daytime. Migratory species in these groups have great difficulty crossing large bodies of water, since thermals only form over land, and these birds cannot maintain active flight for long distances. Mediterranean and other seas present a major obstacle to soaring birds, which must cross at the narrowest points. Massive numbers of large raptors and storks pass through areas such as Gibraltar, Falsterbo, and the Bosphorus at migration times. More common species, such as the European Honey Buzzard, can be counted in hundreds of thousands in autumn. Other barriers, such as mountain ranges, can also cause funnelling, particularly of large diurnal migrants. This is a notable factor in the Central American migratory bottleneck. Batumi bottleneck in the Caucasus is one of the heaviest migratory funnels on earth. Avoiding flying over the Black Sea surface and across high mountains, hundreds of thousands of soaring birds funnel through an area around the city of Batumi, Georgia.[30] It has been suggested [31] that birds of prey (such as Honey Buzzards) which migrate using thermals may benefit from losing 10 to 20% of their weight and that this may explain why they forage less on migration than do smaller birds of prey with more active flight such as Falcons, Hawks and Harriers.

Ruby-throated Hummingbird

Many of the smaller insectivorous birds including the warblers, hummingbirds and flycatchers migrate large distances, usually at night. They land in the morning and may feed for a few days before resuming their migration. The birds are referred to as passage migrants in the regions where they occur for short durations between the origin and destination.[32]

Nocturnal migrants minimize predation, avoid overheating, and feed during the day.[17] One cost of nocturnal migration is the loss of sleep. Migrants may be able to alter their quality of sleep to compensate for the loss.[33]

Short-distance and altitudinal migration

Cedar Waxwing

Many long-distance migrants appear to be genetically programmed to respond to changing day length. Species that move short distances, however, may not need such a timing mechanism, and may move in response to local weather conditions. Thus mountain and moorland breeders, such as Wallcreeper and White-throated Dipper, may move only altitudinally to escape the cold higher ground. Other species such as Merlin and Skylark will move further to the coast or to a more southerly region. Species like the Common Chaffinch are not migratory in Britain, but move south or to Ireland in very cold weather.[citation needed]

Short-distance passerine migrants have two evolutionary origins. Those that have long-distance migrants in the same family, such as the Common Chiffchaff, are species of southern hemisphere origins that have progressively shortened their return migration to stay in the northern hemisphere.[34]

Species that have no long-distance migratory relatives, such as the waxwings, are effectively moving in response to winter weather and the loss of their usual winter food, rather than enhanced breeding opportunities.[35]

Woodland Kingfisher

In the tropics there is little variation in the length of day throughout the year, and it is always warm enough for a food supply (although because of competition, there may not be enough food for every bird). Migration within the tropics has been far less studied than in the temperate zones. It was once assumed that tropical birds were mostly sedentary; however, altitudinal migration and other within-tropics movements appear to be surprisingly common.[citation needed] Many tropical regions have wet and dry seasons, inducing some birds to migrate or wander widely to find food. Indeed, the monsoons of India are preceded by the arrival of the Jacobin Cuckoo, the "harbinger of the monsoon". Other examples include the Woodland Kingfisher of west Africa and many Australian birds.[citation needed]

There are a few species, notably cuckoos, which are genuine long-distance migrants within the tropics. An example is the Lesser Cuckoo, which breeds in India and spends the non-breeding season in Africa. Such examples help make the case that food supplies, not weather per se, drive migration patterns.[citation needed]

Altitudinal migration is common on mountains worldwide, such as in the Himalayas and the Andes.[36] Quite often, altitudinal migration is combined with distance migration; for example, the Himalayan Kashmir Flycatcher and Pied Thrush both move as far south as the highlands of Sri Lanka. Altitudinal migration may even be important to birds living on relatively small islands, such as the Hawaiian Islands, which have high mountains.[citation needed]

Irruptions and dispersal

Sometimes circumstances such as a good breeding season followed by a food source failure the following year lead to irruptions in which large numbers of a species move far beyond the normal range. Bohemian Waxwings well show this unpredictable variation in annual numbers, with five major arrivals in Britain during the nineteenth century, but 18 between the years 1937 and 2000.[35] Red Crossbills too are irruptive, with invasions noted in 1251, 1593, 1757, and 1791.[37]

The temperate zones of the southern continents have extensive arid areas, particularly in Australia and western southern Africa, and weather-driven movements are common but not always predictable. A couple of weeks of heavy rain in one part or another of the usually dry centre of Australia, for example, causes dramatic plant and invertebrate growth, attracting birds from all directions. This can happen at any time of year, and, in any given area, may not happen again for a decade or more, depending on the frequency of El Niño and La Niña periods.[citation needed]

Rainbow Bee-eater

Bird migration is primarily, but not entirely, a Northern Hemisphere phenomenon. In the Southern Hemisphere, seasonal migration tends to be much less obvious. There are several reasons for this.

First, the largely uninterrupted expanses of land mass or ocean tend not to funnel migrations into narrow and obvious pathways, making them less obvious to the human observer. Second, at least for terrestrial birds, climatic regions tend to fade into one another over a long distance rather than be entirely separate: this means that rather than make long trips over unsuitable habitat to reach particular destinations, migrant species can usually travel at a relaxed pace, feeding as they go. Short of banding studies it is often not obvious that the birds seen in any particular locality as the seasons change are in fact different members of the same species passing through, gradually working their way north or south.[citation needed]

Many species do in fact breed in the temperate southern hemisphere regions and winter further north in the tropics. The southern African Greater Striped Swallow, and the Australian Satin Flycatcher, Dollarbird, and Rainbow Bee-eater for example, winters well north of their breeding range.[citation needed]

Physiology and control

The control of migration, its timing and response are genetically controlled and appear to be a primitive trait that is present even in non-migratory species of birds. The ability to navigate and orient themselves during migration is a much more complex phenomenon that may include both endogenous programs as well as learning.[38]

Timing

The primary physiological cue for migration are the changes in the day length. These changes are also related to hormonal changes in the birds.

In the period before migration, many birds display higher activity or Zugunruhe (German: migratory restlessness) as well as physiological changes such as increased fat deposition. The occurrence of Zugunruhe even in cage-raised birds with no environmental cues (e.g. shortening of day and falling temperature) has pointed to the role of circannual endogenous programs in controlling bird migrations. Caged birds display a preferential flight direction that corresponds with the migratory direction they would take in nature, even changing their preferential direction at roughly the same time their wild conspecifics change course.[citation needed]

In species where there is polygyny and with considerable sexual dimorphism, there is a tendency for males to return earlier to the breeding sites than their females. This is termed as protandry.[39][40]

Orientation and navigation

The routes of satellite tagged Bar-tailed Godwits migrating north from New Zealand. This species has the longest known non-stop migration of any species, up to 10,200 km (6,300 mi).

Navigation is based on a variety of senses. Many birds have been shown to use a sun compass. Using the sun for direction involves the need for making compensation based on the time. Navigation has also been shown to be based on a combination of other abilities including the ability to detect magnetic fields (magnetoception), use visual landmarks as well as olfactory cues.[41]

Long distance migrants are believed to disperse as young birds and form attachments to potential breeding sites and to favourite wintering sites. Once the site attachment is made they show high site-fidelity, visiting the same wintering sites year after year.[42]

The ability of birds to navigate during migrations cannot be fully explained by endogenous programming, even with the help of responses to environmental cues. The ability to successfully perform long-distance migrations can probably only be fully explained with an accounting for the cognitive ability of the birds to recognize habitats and form mental maps. Satellite tracking of day migrating raptors such as Ospreys and Honey Buzzards has shown that older individuals are better at making corrections for wind drift.[43]

As the circannual patterns indicate, there is a strong genetic component to migration in terms of timing and route, but this may be modified by environmental influences. An interesting example where a change of migration route has occurred because of such a geographical barrier is the trend for some Blackcaps in central Europe to migrate west and winter in Britain rather than cross the Alps.[citation needed]

Migratory birds may use two electromagnetic tools to find their destinations: one that is entirely innate and another that relies on experience. A young bird on its first migration flies in the correct direction according to the Earth's magnetic field, but does not know how far the journey will be. It does this through a radical pair mechanism whereby chemical reactions in special photo pigments sensitive to long wavelengths are affected by the field. Although this only works during daylight hours, it does not use the position of the sun in any way. At this stage the bird is similar to a boy scout with a compass but no map, until it grows accustomed to the journey and can put its other facilities to use. With experience they learn various landmarks and this "mapping" is done by magnetites in the trigeminal system, which tell the bird how strong the field is. Because birds migrate between northern and southern regions, the magnetic field strengths at different latitudes let it interpret the radical pair mechanism more accurately and let it know when it has reached its destination.[44] More recent research has found a neural connection between the eye and "Cluster N", the part of the forebrain that is active during migrational orientation, suggesting that birds may actually be able to see the magnetic field of the earth.[45][46]

Vagrancy

Migrating birds can lose their way and appear outside their normal ranges. This can be due to flying past their destinations as in the "spring overshoot" in which birds returning to their breeding areas overshoot and end up further north than intended. Reverse migration, where the genetic programming of young birds fails to work properly, can lead to great rarities turning up as vagrants thousands of kilometres out of range. Certain areas, because of their location, have become famous as watchpoints for migrating birds. Examples are the Point Pelee National Park in Canada, and Spurn in England. Drift migration of birds blown off course by the wind can result in "falls" of large numbers of migrants at coastal sites.[citation needed]

A related phenomenon called "abmigration" involves birds from one region joining similar birds from a different breeding region in the common winter grounds and then migrating back along with the new population. This is especially common in some waterfowl, which shift from one flyway to another.[47]

Migration conditioning

It has been possible to teach a migration route to a flock of birds, for example in re-introduction schemes. After a trial with Canada Geese, microlight aircraft were used in the US to teach safe migration routes to reintroduced Whooping Cranes.[48][49]

Adaptations

Birds need to alter their metabolism in order to meet the demands of migration. The storage of energy through the accumulation of fat and the control of sleep in nocturnal migrants require special physiological adaptations. In addition, the feathers of a bird suffer from wear-and-tear and require to be molted. The timing of this molt - usually once a year but sometimes two - varies with some species molting prior to moving to their winter grounds and others molting prior to returning to their breeding grounds.[50][51] Apart from physiological adaptations, migration sometimes requires behavioural changes such as flying in flocks to reduce the energy used in migration or the risk of predation.[52]

Evolutionary and ecological factors

Migration in birds is highly labile and is believed to have developed independently in many avian lineages.[53] While it is agreed that the behavioral and physiological adaptations necessary for migration are under genetic control, some authors have argued that no genetic change is necessary for migratory behavior to develop in a sedentary species because the genetic framework for migratory behavior exists in nearly all avian lineages.[54] This explains the rapid appearance of migratory behavior after the most recent glacial maximum.[55]

Whether a particular species migrates depends on a number of factors. The climate of the breeding area is important, and few species can cope with the harsh winters of inland Canada or northern Eurasia. Thus the partially migratory Blackbird Turdus merula is migratory in Scandinavia, but not in the milder climate of southern Europe. The nature of the staple food is also significant. Most specialist insect eaters outside the tropics are long-distance migrants, and have little choice but to head south in winter.[citation needed]

Sometimes the factors are finely balanced. The Whinchat Saxicola rubetra of Europe and the Siberian Stonechat Saxicola maura of Asia are long-distance migrants wintering in the tropics, whereas their close relative, the European Stonechat Saxicola rubicola is a resident bird in most of its range, and moves only short distances from the colder north and east. A possible factor here is that the resident species can often raise an extra brood.[citation needed]

Recent research suggests that long-distance passerine migrants are of South American and African, rather than northern hemisphere, evolutionary origins. They are effectively southern species coming north to breed rather than northern species going south to winter.[citation needed]

Theoretical analyses, summarized by Alerstam (2001), show that detours that increase flight distance by up to 20% will often be adaptive on aerodynamic grounds - a bird that loads itself with food to cross a long barrier flies less efficiently. However some species show circuitous migratory routes that reflect historical range expansions and are far from optimal in ecological terms. An example is the migration of continental populations of Swainson's Thrush, which fly far east across North America before turning south via Florida to reach northern South America; this route is believed to be the consequence of a range expansion that occurred about 10,000 years ago. Detours may also be caused by differential wind conditions, predation risk, or other factors.[citation needed]

Climate change

Large scale climatic changes, as have been experienced in the past, are expected to have an effect on the timing of migration. Studies have shown a variety of effects including timing changes in migration, breeding[56] as well as population variations.[57][58]

Ecological effects

The migration of birds also aids the movement of other species, including those of ectoparasites such as ticks and lice,[59] which in turn may carry micro-organisms including those of concern to human health. Considerable interest has been taken due to the global spread of avian influenza, however migrant birds have not been found to be a special risk, with import of pet and domestic birds being a greater threat.[60] Some viruses that are maintained in birds without lethal effects, such as the West Nile Virus may however be spread by migrating birds.[61] Birds may also have a role in the dispersal of propagules of plants and plankton.[62][63]

Some predators take advantage of the concentration of birds during migration. Greater Noctule bats feed on nocturnal migrating passerines.[5] Some birds of prey specialize on migrating waders.[64]

Study techniques

Radars for monitoring bird migration. Kihnu, Estonia.

Early studies on the timing of migration began in 1749 in Finland, with Johannes Leche of Turku collecting the dates of arrivals of spring migrants.[65]

Bird migration routes have been studied by a variety of techniques including the oldest, marking. Swans have been marked with a nick on the beak since about 1560 in England. Scientific ringing was pioneered by Mortensen in 1899.[66] Other techniques include radar and satellite tracking.[citation needed]

Stable isotopes of hydrogen, oxygen, carbon, nitrogen, and sulphur can establish avian migratory connectivity between wintering sites and breeding grounds. Stable isotopic methods to establish migratory linkage rely on spatial isotopic differences in bird diet that are incorporated into inert tissues like feathers, or into growing tissues such as claws and muscle or blood.[67][68]

An approach to identify migration intensity makes use of upward pointing microphones to record the nocturnal contact calls of flocks flying overhead. These are then analyzed in a laboratory to measure time, frequency and species.[69]

Emlen funnel

An older technique to quantify migration involves observing the face of the moon towards full moon and counting the silhouettes of flocks of birds as they fly at night.[70][71]

Orientation behaviour studies have been traditionally carried out using variants of a setup known as the Emlen funnel, which consists of a circular cage with the top covered by glass or wire-screen so that either the sky is visible or the setup is placed in a planetarium or with other controls on environmental cues. The orientation behaviour of the bird inside the cage is studied quantitatively using the distribution of marks that the bird leaves on the walls of the cage.[72] Other approaches used in pigeon homing studies make use of the direction in which the bird vanishes on the horizon.

Threats and conservation

Migration routes and countries with illegal hunting in Europe

Human activities have threatened many migratory bird species. The distances involved in bird migration mean that they often cross political boundaries of countries and conservation measures require international cooperation. Several international treaties have been signed to protect migratory species including the Migratory Bird Treaty Act of 1918 of the US.[73] and the African-Eurasian Migratory Waterbird Agreement[74]

The concentration of birds during migration can put species at risk. Some spectacular migrants have already gone extinct, the most notable being the Passenger Pigeon (Ectopistes migratorius). During migration the enormous flocks were a mile (1.6 km) wide, darkening the sky and 300 miles (480 km) long, taking several days to pass.[75]

Other significant areas include stop-over sites between the wintering and breeding territories.[76] A capture-recapture study of passerine migrants with high fidelity for breeding and wintering sites did not show similar strict association with stop-over sites.[77]

Hunting along the migratory route can also take a heavy toll. The populations of Siberian Cranes that wintered in India declined due to hunting along the route, particularly in Afghanistan and Central Asia. Birds were last seen in their favourite wintering grounds in Keoladeo National Park in 2002.[78]

Structures such as power lines, wind farms and offshore oil-rigs have also been known to affect migratory birds.[79] Habitat destruction by land use changes is the biggest threat, and shallow wetlands that are stopover and wintering sites for migratory birds are particularly threatened by draining and reclamation for human use.

See also

References

  1. ^ Peter Berthold, Hans-Günther Bauer, Valerie Westhead (2001). Bird Migration: A General Survey. Oxford: Oxford University Press. ISBN 0-19-850787-9.{{cite book}}: CS1 maint: multiple names: authors list (link)
  2. ^ Sekercioglu, C.H. (2007). "Conservation ecology: area trumps mobility in fragment bird extinctions". Current Biology. 17 (8): 283–286.
  3. ^ a b c d e Newton, I. (2008). The Migration Ecology of Birds. Elselvier. ISBN 978-0-12-517367-4.
  4. ^ Dondini, G., Vergari, S. (2000). "Carnivory in the greater noctule bat (Nyctalus lasiopterus) in Italy". Journal of Zoology. 251 (2): 233–236. doi:10.1111/j.1469-7998.2000.tb00606.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. ^ a b Popa-Lisseanu, A. G., Delgado-Huertas, A., Forero, M. G., Rodriguez, A., Arlettaz, R. & Ibanez, C. (2007). Rands, Sean (ed.). "Bats' Conquest of a Formidable Foraging Niche: The Myriads of Nocturnally Migrating Songbirds". PLoS ONE. 2 (2): e205. doi:10.1371/journal.pone.0000205. PMC 1784064. PMID 17299585.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  6. ^ Ibáñez, C., Juste, J., García-Mudarra, J. L., Agirre-Mendi, P. T. (2001). "Bat predation on nocturnally migrating birds". PNAS. 98 (17): 9700–9702. doi:10.1073/pnas.171140598. PMC 55515. PMID 11493689.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  7. ^ Chan K (2001). "Partial migration in Australian landbirds: a review". Emu. 101 (4): 281–292. doi:10.1071/MU00034.
  8. ^ Boland, J. M. (1990). "Leapfrog migration in North American shorebirds: intra- and interspecific examples" (PDF). The Condor. 92 (2): 284–290. doi:10.2307/1368226. JSTOR 1368226.
  9. ^ Panov, Ilya N (2011). "Overlap between moult and autumn migration in passerines in northern taiga zone of Eastern Fennoscandia" (PDF). Avian Ecology and Behaviour. 19: 33–64.
  10. ^ Hummel D., Beukenberg M. (1989). "Aerodynamische Interferenzeffekte beim Formationsfl ug von Vogeln". J. Ornithol. 130: 15–24. doi:10.1007/BF01647158.
  11. ^ Cutts, C. J. & J R Speakman (1994). "Energy savings in formation flight of Pink-footed Geese" (PDF). J. Exp. Biol. 189 (1): 251–261. PMID 9317742.
  12. ^ Geroudet, P. (1954). "Des oiseaux migrateurs trouvés sur la glacier de Khumbu dans l'Himalaya". Nos Oiseaux. 22: 254.
  13. ^ Swan, L. W. (1970). "Goose of the Himalayas". Nat. Hist. 79 (10): 68–75.
  14. ^ Dorst, J. (1963). The migration of birds. Houghton Mifflin Co., Boston. p. 476.
  15. ^ Eastwood, E. & G. C. Rider. (1965). "Some radar measurements of the altitude of bird flight". British Birds. 58: 393–426.
  16. ^ Williams, G. G. (1950). "Weather and spring migration". Auk. 67: 52–65.
  17. ^ a b c Lincoln, F. C. (1979). Migration of Birds. Fish and Wildlife Service. Circular 16. {{cite book}}: Italic or bold markup not allowed in: |publisher= (help)
  18. ^ Cocker, Mark; Mabey, Richard (2005). Birds Britannica. Chatto & Windus. p. 315. ISBN 0-7011-6907-9.{{cite book}}: CS1 maint: multiple names: authors list (link)
  19. ^ a b White, 1898. pp. 27–28
  20. ^ White, 1898. pp. 161–162
  21. ^ Bewick, 1797. p. xvii
  22. ^ Bewick, 1797. p. 300
  23. ^ Bewick, 1797. pp. 302–303
  24. ^ Dark-Eyed Junco
  25. ^ American Goldfinch
  26. ^ Sprague, A. J., D. J. Hamilton, and A. W. Diamond (2008). "Site Safety and Food Affect Movements of Semipalmated Sandpipers (Calidris pusilla) Migrating Through the Upper Bay of Fundy" (PDF). Avian Conservation and Ecology. 3 (2). Retrieved 8 May 2013.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  27. ^ Kathleen E. Clark, Lawrence J. Niles and Joanna Burger. "Abundance and Distribution of Migrant Shorebirds in Delaware Bay" (PDF). The Condor (95): 694–705. Retrieved 8 May 2013.
  28. ^ Gill, Robert E. Jr., Theunis Piersma, Gary Hufford, Rene Servranckx, Adrian Riegen (2005). "Crossing the ultimate ecological barrier: evidence for an 11,000 km-long nonstop flight from Alaska to New Zealand and Eastern Australia by Bar-tailed Godwits". The Condor. 107 (1): 1–20. doi:10.1650/7613.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  29. ^ Anon (18 April 2002). "Oldest bird clocks 5 million miles". CNN.com. Retrieved 31 March 2013.
  30. ^ Maanen, E. van; Goradze, I.; Gavashelishvili, A.; Goradze, R. (2001). "Opinion: Trapping and hunting of migratory raptors in western Georgia". Bird Conservation International. 11 (2): 77–92. doi:10.1017/S095927090100017X.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  31. ^ Gensbol, B; (1984) Collins Guide to the Birds of Prey of Britain and Europe, p.28
  32. ^ Schmaljohann, Heiko, Felix Liechti and Bruno Bruderer (2007). "Songbird migration across the Sahara: the non-stop hypothesis rejected!". Proceedings of the Royal Society B. 274 (1610): 735–739. doi:10.1098/rspb.2006.0011. PMC 2197203. PMID 17254999.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  33. ^ Rattenborg, N.C., Mandt, B.H., Obermeyer, W.H., Winsauer, P.J., Huber, R. (2004). "Migratory Sleeplessness in the White-Crowned Sparrow (Zonotrichia leucophrys gambelii)". PLoS Biol. 2 (7): e212. doi:10.1371/journal.pbio.0020212. PMC 449897. PMID 15252455.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  34. ^ Cocker, 2005. p. 378
  35. ^ a b Cocker, 2005. p. 326
  36. ^ Kreft, Stefan (23 June 2004). "The Fourth Dimension: An Overview of Altitudinal Migration" (PDF). 25th Annual Bonn Convention, Berlin. Retrieved 27 March 2013.
  37. ^ Cocker, 2005. p. 455
  38. ^ Helm B, Gwinner E (2006). "Migratory Restlessness in an Equatorial Nonmigratory Bird". PLoS Biol. 4 (4): e110. doi:10.1371/journal.pbio.0040110. PMC 1420642. PMID 16555925.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  39. ^ Diego Rubolini, Fernando Spina and Nicola Saino (2004). "Protandry and sexual dimorphism in trans-Saharan migratory birds". Behavioral Ecology. 15 (4): 592–601. doi:10.1093/beheco/arh048.
  40. ^ Edwards, Darryl B.; Forbes, Mark R. (2007). "Absence of protandry in the spring migration of a population of Song Sparrows Melospiza melodia". Ibis. 149 (4): 715–720. doi:10.1111/j.1474-919X.2007.00692.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  41. ^ Walraff, H. G. (2005). Avian Navigation: Pigeon Homing as a Paradigm. Springer.
  42. ^ Ketterson, E.D. and V. Nolan Jr. (1990). "Site attachment and site fidelity in migratory birds: experimental evidence from the field and analogies from neurobiology.". In E. Gwinner (ed.). Bird Migration. Springer Verlag. pp. 117–129. Archived from the original (PDF) on 2010-02-10.
  43. ^ Thorup, Kasper, Thomas Alerstam, Mikael Hake and Nils Kjelle (2003). "Bird orientation: compensation for wind drift in migrating raptors is age dependent". Proceedings of the Royal Society B. 270 (Suppl 1): S8–S11. doi:10.1098/rsbl.2003.0014. PMC 1698035. PMID 12952622.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  44. ^ Wiltschko, W., U. Munro, H. Ford & R. Wiltschko (2006). "Bird navigation: what type of information does the magnetite-based receptor provide?". Proceedings of the Royal Society B. 273 (1603): 2815–20. doi:10.1098/rspb.2006.3651. PMC 1664630. PMID 17015316.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  45. ^ Heyers D, Manns M, Luksch H, Güntürkün O, Mouritsen H (2007). Iwaniuk, Andrew (ed.). "A Visual Pathway Links Brain Structures Active during Magnetic Compass Orientation in Migratory Birds". PLoS ONE. 2 (9): e937. doi:10.1371/journal.pone.0000937. PMC 1976598. PMID 17895978.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  46. ^ Deutschlander, ME, Phillips, JB, Borland, SC (1999). "The case for light-dependent magnetic orientation in animals" (PDF). J.Exp. Biol. 202 (8): 891–908. PMID 10085262.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  47. ^ Guillemain, M., Sadoul, N. and Simon, G. (2005), European flyway permeability and abmigration in Teal Anas crecca, an analysis based on ringing recoveries. Ibis, 147: 688–696. doi: 10.1111/j.1474-919X.2005.00446.x
  48. ^ "Operation migration".
  49. ^ "Wisconsin Whooping Crane Management Plan" (PDF). Wisconsin Department of Natural Resources. 6 December 2006.
  50. ^ Rohwer S; Butler LK & DR Froehlich (2005). "Ecology and Demography of East-West Differences in Molt Scheduling of Neotropical Migrant Passerines". In Greenberg R & Marra PP (ed.). Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press. p. 94. ISBN 0-8018-8107-2.{{cite book}}: CS1 maint: multiple names: authors list (link)
  51. ^ Hedenström, A (2008). "Adaptations to migration in birds: behavioural strategies, morphology and scaling effects". Philosophical Transactions of the Royal Society B. 363 (1490): 287–299. doi:10.1098/rstb.2007.2140. PMC 2606751. PMID 17638691.
  52. ^ Weber, Jean-Michel (2009). "The physiology of long-distance migration: extending the limits of endurance metabolism". J. Exp. Biol. 212 (Pt 5): 593–597. doi:10.1242/jeb.015024. PMID 19218508.
  53. ^ F. Pulido. (2007). "The genetics and evolution of avian migration". BioScience. 57: 165–174.
  54. ^ J. Rappole, B. Helm, M. Ramos. (2003). "An integrative framework for understanding the origin and evolution of avian migration". Journal of Avian Biology. 34: 125.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  55. ^ B. Mila, T. Smith, R. Wayne. (2006). "Postglacial population expansion drives the evolution of long-distance avian migration in a songbird". Evolution. 60: 2403–2409.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  56. ^ Jenni L. & Kery M. (2003). "Timing of autumn bird migration under climate change: advances in long-distance migrants, delays in short-distance migrants". Proceedings of the Royal Society B. 270: 1467.
  57. ^ Both, Christiaan (2006-05-04). "Climate change and population declines in a long-distance migratory bird". Nature. 441 (7089): 81–83. doi:10.1038/nature04539. ISSN 0028-0836. PMID 16672969. {{cite journal}}: Unknown parameter |coauthors= ignored (|author= suggested) (help)
  58. ^ Wormworth, J. & K. Mallon (2006). Bird Species and Climate Change: The Global Status Report version 1.0. WWF.
  59. ^ Smith RP Jr, Rand PW, Lacombe EH, Morris SR, Holmes DW, Caporale DA (1996). "Role of bird migration in the long-distance dispersal of Ixodes dammini, the vector of Lyme disease". J. Infect. Dis. 174 (1): 221–4. PMID 8656000.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  60. ^ Rappole, J.H., Hubálek, Zdenek (2006). "Birds and Influenza H5N1 Virus Movement to and within North America". Emerging Infectious Diseases. 12: 10. hdl:10088/875.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  61. ^ Rappole, J.H., Derrickson, S.R., Hubalek, Z. (2000). "Migratory birds and spread of West Nile virus in the Western Hemisphere". Emerging Infectious Diseases. 6 (4): 319–328. doi:10.3201/eid0604.000401. PMC 2640881. PMID 10905964. hdl:10088/364.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  62. ^ Figuerola, O. and A. J. Green (2002). "Dispersal of aquatic organisms by waterbirds: a review of past research and priorities for future studies". Freshwater Biology. 47 (3): 483–494. doi:10.1046/j.1365-2427.2002.00829.x.
  63. ^ Cruden, R. W. (1966). "Birds as Agents of Long-Distance Dispersal for Disjunct Plant Groups of the Temperate Western Hemisphere". Evolution. 20 (4). Evolution, Vol. 20, No. 4: 517–532. doi:10.2307/2406587. JSTOR 2406587.
  64. ^ Ronald C. Ydenberg, Robert W. Butler, David B. Lank, Barry D. Smith and John Ireland (2004). "Western sandpipers have altered migration tactics as peregrine falcon populations have recovered" (PDF). Proceedings of the Royal Society B. 271 (1545): 1263–1269 1263. doi:10.1098/rspb.2004.2713. PMC 1691718. PMID 15306350.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  65. ^ Greenwood, Jeremy J. D. (2007). "Citizens, science and bird conservation". J. Ornithol . 148 (Suppl 1): S77–S124. doi:10.1007/s10336-007-0239-9.
  66. ^ Spencer, R. (1985) Marking. In: Campbell. B. & Lack, E. 1985. A dictionary of birds. British Ornithologists' Union. London, pp. 338–341.
  67. ^ Keith Hobson, Leonard Wassenaar (1997). "Linking breeding and wintering grounds of neotropical migrant songbirds using stable hydrogen isotopic analysis of feathers". Oecologia. 109: 142–148. doi:10.1007/s004420050068.
  68. ^ Gabriel Bowen, Leonard Wassenaar, Keith Hobson (2005). "Global application of stable hydrogen and oxygen isotopes to wildlife forensics". Oecologia. 143 (3): 337–348. doi:10.1007/s00442-004-1813-y. PMID 15726429.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  69. ^ Farnsworth, A., Gauthreaux, S.A., and van Blaricom, D.; Gauthreaux (2004). "A comparison of nocturnal call counts of migrating birds and reflectivity measurements on Doppler radar" (PDF). Journal of Avian Biology. 35 (4): 365–369. doi:10.1111/j.0908-8857.2004.03180.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  70. ^ Liechti, F. (1996). Instructions to count nocturnal bird migration by watching the full moon. Schweizerische Vogelwarte, CH-6204 Sempach, Switzerland.
  71. ^ Lowery, G.H. (1951). "A quantitative study of the nocturnal migration of birds". University Kan. Pub. Mus. Nat. Hist. 3: 361–472.
  72. ^ Emlen, S. T. and Emlen, J. T. (1966). "A technique for recording migratory orientation of captive birds". Auk. 83: 361–367.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  73. ^ "Migratory bird Treaty 16 USC 703-711; 40 Stat. 755". Legal Information Institute (LII). Cornell Law School.
  74. ^ "African-Eurasian Migratory Waterbird Agreement".
  75. ^ "The Passenger Pigeon". Smithsonian Institution. Retrieved 2013-05-24.
  76. ^ Shimazaki, Hiroto; Masayuki Tamura and Hiroyoshi Higuchi (2004). "Migration routes and important stopover sites of endangered oriental white storks (Ciconia boyciana) as revealed by satellite tracking" (PDF). Mem Natl Inst. Polar Res., Spec. Issue. 58: 162–178.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  77. ^ Catry, P., Encarnacao, V., Araujo, A., Fearon, P., Fearon, A., Armelin, M. & Delaloye, P. (2004). "Are long-distance migrant passerines faithful to their stopover sites?" (PDF). Journal of Avian Biology. 35 (2): 170–181. doi:10.1111/j.0908-8857.2004.03112.x.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  78. ^ "Siberian Crane fact sheet".
  79. ^ "Fish and Wildlife Service- Bird Mortality Fact sheet" (PDF).

Further reading

  • Alerstam, Thomas (2001). "Detours in bird migration" (PDF). Journal of Theoretical Biology. 209: 319–331.
  • Berthold, Peter (2001) Bird Migration: A General Survey. Second Edition. Oxford University Press. ISBN 0-19-850787-9
  • Bewick, Thomas (1797–1804). History of British Birds (1847 ed.). Newcastle: Beilby and Bewick.
  • Dingle, Hugh. Migration: The Biology of Life on The Move. Oxford University Press, 1996.
  • Hobson, Keith and Wassenaar, Leonard (2008) Tracking Animal Migration with Stable Isotopes. Academic Press. ISBN 978-0-12-373867-7
  • Weidensaul, Scott. Living On the Wind: Across the Hemisphere With Migratory Birds. Douglas & McIntyre, 1999.
  • Dedicated issue of Philosophical Transactions B on Adaptation to the Annual Cycle.
  • White, Gilbert (1898 (first published 1789)). The Natural History of Selborne. Walter Scott. {{cite book}}: Check date values in: |year= (help)

External links

Template:Link GA Template:Link FA Template:Link FA